advanced thermodynamics for engineers

393
Advanced Thermodynamics for Engineers Desmond E Winterbone FEng, BSc, PhD, DSc, FIMechE, MSAE Thermodynamics and Fluid Mechanics Division Department of Mechanical Engineering UMIST A member of the Hodder Headline Group LONDON SYDNEY AUCKLAND Copublished in North, Central and South America by John Wiley & Sons, he., New York Toronto

Upload: amit-mittal

Post on 06-Mar-2015

819 views

Category:

Documents


22 download

TRANSCRIPT

Page 1: Advanced Thermodynamics for Engineers

Advanced Thermodynamics for Engineers

Desmond E Winterbone

FEng, BSc, PhD, DSc, FIMechE, MSAE

Thermodynamics and Fluid Mechanics Division Department of Mechanical Engineering UMIST

A member of the Hodder Headline Group

LONDON SYDNEY AUCKLAND Copublished in North, Central and South America by John Wiley & Sons, he., New York Toronto

Page 2: Advanced Thermodynamics for Engineers

First published in Great Britain 1997 by Arnold, a member of the Hodder Headline Group, 338 Euston Road, London NW1 3BH

Copublished in North, Central and South America by John Wiley & Sons, Inc., 605 Third Avenue, New York, NY 10158-0012

0 1997 D E Winterbone

All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means, electronically or mechanically, including photocopying, recording or any information storage or retrieval system, without either prior permission in writing from the publisher or a licence permitting restricted copying. In the United Kingdom such licences are issued by the Copyright Licensing Agency: 90 Tottenham Court Road, London WIP 9HE.

Whilst the advice and information in this book is believed to be true and accurate at the date of going to press, neither the author nor the publisher can accept any legal responsibility or liability for any errors or omissions that may be made.

British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library

Library of Congress Cataloging-in-Publication Data A catalog record for this book is available from the Library of Congress

ISBN 0 340 67699 X (pb) 0 470 23718 X (Wiley)

Typeset in 10/12 pt Times by Mathematical Composition Setters Ltd, Salisbury, Wilts

Printed and bound in Great Britain by J W Anowsmith Ltd, Bristol

Page 3: Advanced Thermodynamics for Engineers

Preface

When reviewing, or contemplating writing, a textbook on engineering thermodynamics, it is necessary to ask what does this book offer that is not already available? The author has taught thermodynamics to mechanical engineering students, at both undergraduate and post-graduate level, for 25 years and has found that the existing texts cover very adequately the basic theories of the subject. However, by the final years of a course, and at post-graduate level, the material which is presented is very much influenced by the lecturer, and here it is less easy to find one book that covers all the syllabus in the required manner. This book attempts to answer that need, for the author at least.

The engineer is essentially concerned with manufacturing devices to enable tasks to be preformed cost effectively and efficiently. Engineering has produced a new generation of automatic ‘slaves’ which enable those in the developed countries to maintain their lifestyle by the consumption of fuels rather than by manual labour. The developing countries still rely to a large extent on ‘manpower’, but the pace of development is such that the whole world wishes to have the machines and quality of life which we, in the developed countries, take for granted: this is a major challenge to the engineer, and particularly the thermodynamicist. The reason why the thermodynamicist plays a key role in this scenario is because the methods of converting any form of energy into power is the domain of thermodynamics: all of these processes obey the four laws of thermodynamics, and their efficiency is controlled by the Second Law. The emphasis of the early years of an undergraduate course is on the First Law of thermodynamics, which is simply the conservation of energy; the First Law does not give any information on the quality of the energy. It is the hope of the author that this text will introduce the concept of the quality of energy and help future engineers use our resources more efficiently. Ironically, some of the largest demands for energy may come from cooling (e.g. refrigeration and air- conditioning) as the developing countries in the tropical regions become wealthier - this might require a more basic way of considering energy utiiisation than that emphasised in current thermodynamic texts. This book attempts to introduce basic concepts which should apply over the whole range of new technologies covered by engineering thermodynamics. It considers new approaches to cycles, which enable their irreversibility to be taken into account; a detailed study of combustion to show how the chemical energy in a fuel is converted into thermal energy and emissions; an analysis of fuel cells to give an understanding of the direct conversion of chemical energy to electrical power; a detailed study of property relationships to enable more sophisticated analyses to be made of both

Page 4: Advanced Thermodynamics for Engineers

x Preface

high and low temperature plant; and irreversible thermodynamics, whose principles might hold a key to new ways of efficiently converting energy to power (e.g. solar energy, fuel cells).

The great advances in the understanding and teaching of thermodynamics came rapidly towards the end of the 19th century, and it was not until the 1940s that these were embodied in thermodynamics textbooks for mechanical engineers. Some of the approaches used in teaching thermodynamics st i l l contain the assumptions embodied in the theories of heat engines without explicitly recognising the limitations they impose. It was the desire to remove some of these shortcomings, together with an increasing interest in what limits the efficiency of thermodynamic devices, that led the author down the path that has culminated in this text.

I am sti l l a strong believer in the pedagogical necessity of introducing thermodynamics through the traditional route of the Zeroth, First, Second and Third Laws, rather than attempting to use the Single-Axiom Theorem of Hatsopoulos and Keenan, or The Law of Stable Equilibrium of Haywood. While both these approaches enable thermodynamics to be developed in a logical manner, and limit the reliance on cyclic processes, their understanding benefits from years of experience - the one thing students are lacking. I have structured this book on the conventional method of developing the subject. The other dilemma in developing an advanced level text is whether to introduce a significant amount of statistical thermodynamics; since this subject is related to the particulate nature of matter, and most engineers deal with systems far from regions where molecular motion dominates the processes, the majority of the book is based on equilibrium ther- modynamics; which concentrates on the macroscopic nature of systems. A few examples of statistical thermodynamics are introduced to demonstrate certain forms of behaviour, but a full understanding of the subject is not a requirement of the text.

The book contains 17 chapters and, while this might seem an excessive number, these are of a size where they can be readily incorporated into a degree course with a modular structure. Many such courses will be based on two hours lecturing per week, and this means that most of the chapters can be presented in a single week. Worked examples are included in most of the chapters to illustrate the concepts being propounded, and the chapters are followed by exercises. Some of these have been developed from texts which are now not available (e.g. Benson, Haywood) and others are based on examination questions. Solutions are provided for all the questions. The properties of gases have been derived from polynomial coefficients published by Benson: all the parameters quoted have been evaluated by the author using these coefficients and equations published in the text - this means that all the values are self-consistent, which is not the case in all texts. Some of the combustion questions have been solved using computer programs developed at UMIST, and these are all based on these gas property polynomials. If the reader uses other data, e.g. JANAF tables, the solutions obtained might differ slightly from those quoted.

Engineering thermodynamics is basically equilibrium thermodynamics, although for the first two years of the conventional undergraduate course these words are used but not often defined. Much of the thermodynamics done in the early years of a course also relies heavily on reversibilio, without explicit consideration of the effects of irreversibility. Yet, if the performance of thermodynamic devices is to be improved, it is the irreversibility that must be tackled. This book introduces the effects of irreversibility through considerations of availability (exergy), and the concept of the endoreversible engine. The thermal efficiency is related to that of an ideal cycle by the rational efficiency - to demonstrate how closely the performance of an engine approaches that of a reversible one. It is also

Page 5: Advanced Thermodynamics for Engineers

Preface xi

shown that the Camot efficiency is a very artificial yardstick against which to compare real engines: the internal and external reversibilities imposed by the cycle mean that it produces zero power at the maximum achievable efficiency. The approach by CuIZon and Ahlbom to define the efficiency of an endoreversible engine producing maximum power output is introduced: this shows the effect of extern1 irreversibility. This analysis also introduces the concept of entropy generation in a manner readily understandable by the engineec this concept is the comerstone of the theories of irreversible thennodynamics which are at the end of the text.

Whilst the laws of thermodynamics can be developed in isolation from consideration of the property relationships of the system under consideration, it is these relationships that enable the equations to be closed. Most undergraduate texts are based on the evaluation of the fluid properties from the simple perfect gas law, or from tables and charts. While this approach enables typical engineering problems to be solved, it does not give much insight into some of the phenomena which can happen under certain circumstances. For example, is the specific heat at constant volume a function of temperature alone for gases in certain regions of the state diagram? Also, why is the assumption of constant stagnation, or even static, temperature valid for flow of a perfect gas through a throttle, but never for steam? An understanding of these effects can be obtained by examination of the more complex equations of state. This immediately enables methods of gas liquefaction to be introduced.

An important area of enginee~g thermodynamics is the combustion of hydrocarbon fuels. These fuels have formed the driving force for the improvement of living standards which has been seen over the last century, but they are presumably finite, and are producing levels of pollution that are a constant challenge to engineers. At present, there is the threat of global warming due to the build-up of carbon dioxide in the atmosphere: this requires more efficient engines to be produced, or for the carbon-hydrogen ratio in fuels to be reduced. Both of these are major challenges, and while California can legislate for the Zero Emissions Vehicle (ZEV) this might not be a worldwide solution. It is said that the ZEV is an electric car running in Los Angeles on power produced in Arizona! - obviously a case of exporting pollution rather than reducing it. The real challenge is not what is happening in the West, although the energy consumption of the USA is prodigious, but how can the aspirations of the East be met. The combustion technologies developed today will be necessary to enable the Newly Industrialised Countries (NICs) to approach the level of energy consumption we enjoy. The section on combustion goes further than many general textbooks in an attempt to show the underlying general principles that affect combustion, and it introduces the interaction between thermodynamics and fluid mechanics which is so important to achieving clean and efficient combustion. The final chapter introduces the thermodynamic principles of fuel cells, which enable the direct conversion of the Gibbs energy in the fuel to electrical power. Obviously the fuel cell could be a major contributor to the production of 'clean' energy and is a goal for which it is worth aiming.

Finally, a section is included on irreversible thermodynamics. This is there partly as an intellectual challenge to the reader, but also because it infroduces concepts that might gain more importance in assessing the performance of advanced forms of energy conversion. For example, although the fuel cell is basically a device for converting the Gibbs energy of the reactants into electrical energy, is its efficiency compromised by the thermodynamics of the steady state that are taking place in the cell? Also, will photo-voltaic devices be limited by phenomena considered by irreversible thermodynamics?

Page 6: Advanced Thermodynamics for Engineers

xii Preface

I have taken the generous advice of Dr Joe Lee, a colleague in the Department of Chemistry, UMIST, and modified some of the wording of the original text to bring it in line with more modem chemical phraseology. I have replaced the titles Gibbs free energy and Helmholtz free energy by Gibbs and Helmholtz energy respectively: this should not cause any problems and is more logical than including the word ‘free’. I have bowed, with some reservations, to using the internationally agreed spelling sulfur, which again should not cause problems. Perhaps the most difficult concept for engineers will be the replacement of the terms ‘mol’ and ‘kmol’ by the term ‘amount of substance’. This has been common practice in chemistry for many years, and separates the general concept of a quantity of matter from the units of that quantity. For example, it is common to talk of a mass of substance without defining whether it is in grams, kilograms, pounds, or whatever system of units is appropriate. The use of the phrase ‘amount of substance’ has the same generalising effect when dealing with quantities based on molecular equivalences. The term mol will still be retained as the adjective and hence molar enthalpy is the enthalpy per unit amount of substance in the appropriate units (e.g. kJ/mol, kJ/kmol, Btu/lb-mol, etc).

I would like to acknowledge all those who have helped and encouraged the writing of this text. First, I would like to acknowledge the influence of all those who attempted to teach me thermodynamics; and then those who encouraged me to teach the subject, in particular Jim Picken, Frank Wallace and Rowland Benson. In addition, I would like to acknowledge the encouragement to develop the material on combustion which I received from Roger Green during an Erskine Fellowship at the University of Canterbury, New Zealand. Secondly, I would like to thank those who have helped in the production of this book by reading the text or preparing some of the material. Amongst these are Ed Moses, Marcus Davies, Poh Sung Loh, Joe Lee, Richard Pearson and John Horlock; whilst they have read parts of the text and provided their comments, the responsibility for the accuracy of the book lies entirely in my hands. I would also like to acknowledge my secretary, Mrs P Shepherd, who did some of the typing of the original notes. Finally, I must thank my wife, Veronica, for putting up with lack of maintenance in the house and garden, and many evenings spent alone while I concentrated on this work.

D E Winterbone

Page 7: Advanced Thermodynamics for Engineers

Ad 0

0

0

Q)

n

+

L 3

0 3

+J

L

G

Page 8: Advanced Thermodynamics for Engineers

Symbols

a a a Of

ai A A b b B CP

C d D e E E E EO E , f F F

activity coefficient specific non-flow availability coefficient in van der Waals' equation specific flow availability enthalpy coefficient for gas properties non- flow availability area specific exergy coefficient in van der Waals' equation exergy specific heat capacity at constant pressure (sometimes abbreviated to specific heat at constant pressure) mean value of specific heat capacity at constant pressure over a range of temperatures molar specific heat capacity at constant pressure (Le. specific heat capacity at constant pressure based on rnols) specific heat capacity at constant volume (sometimes abbreviated to specific heat at constant volume) molar specific heat capacity at constant volume (i.e. specific heat capacity at constant volume based on rnols) conductivity for heat flow into engines increment in - usually used for definite integral, e.g. property, etc mass diffusivity specific internal energy internal energy activation energy electromotive force of a cell (emf) standard emf of a cell open circuit voltage specific Helmholtz energy (specific Helmholtz function) Helmholtz energy (Helmholtz function) force

Page 9: Advanced Thermodynamics for Engineers

mi Symbols

F

g go g G h

h H I I J J I

JQ Js k k k

k K

ho

ks

KP 1 1, L Le m mw n n

n

P Pi

Po

4 41

N

pr

Q Q* QP Qb Q" Q: R R

Faraday constant (charge carried by kmol of unit positive valency [96 487 kC/kmol]) specific Gibbs energy (specific Gibbs function) specific Gibbs energy at datum temperature (or absolute zero) acceleration due to gravity Gibbs energy (Gibbs function) specific enthalpy specific enthalpy at datum temperature (or absolute zero) height enthalpy irreversibility electrical current thermodynamic velocity, or flow electrical flow rate heat flow rate entropy flow rate isothermal compressibility, isothermal bulk modulus thermal conductivity Boltzmann constant [ 1.380 62 x 10 -23 J/K] adiabatic, or isentropic, compressibility rate of reaction Karlovitz number [(u'/lT)(d,/u,)] equilibrium constant length Taylor microscale coefficient relating thermodynamic force and velocity Lewis number [rl/pc,D] mass molecular weight polytropic index amount of substance, chemical amount (sometimes referred to as number of mols) reaction order Avogadro constant (6.023 x loz6 kmol-') pressure partial pressure of component i datum pressure (often 1 bar or 1 atmosphere) Prandtl number [ c#/ k] specific heat (energy) transfer electrical charge heat (energy) transfer heat of transport enthalpy of reaction (energy of reaction at constant pressure) calorific value (at constant pressure) = - Qp internal energy of reaction (energy of reaction at constant volume) calorific value (at constant volume) = - Q, specific gas constant rate of formation, rate of reaction

Page 10: Advanced Thermodynamics for Engineers

Symbols xvii

R R R Re

SO S

t t T

UO U’

UI

ut

U U

V V V

S

S*

U

V

V P W $ W X

X

X

X Y 2

radical electrical resistance universal gas constant Reynolds number [pvZ/p I specific entropy specific entropy at datum temperature (or absolute zero) entropy entropy of transport temperature on discontinuous scale time temperature on absolute scale (thermodynamic temperature) specific intrinsic internal energy specific intrinsic internal energy at datum temperature (or absolute zero) turbulence intensity laminar burning velocity turbulent burning velocity intrinsic internal energy overall heat transfer coefficient specific volume volume velocity voltage mean piston speed (for reciprocating engines) specific work maximum specific work work dryness fraction (quality) molar fraction distance thermodynamic force mass fraction valency

Greek characters

degree of dissociation

branching multiplication coefficient molecular thermal diffusivity crank angle (in internal combustion engines) coefficient of thermal expansion

increment in - usually used for indefinite integral, e.g. work (W), heat (Q)

increment in - usually used for indefinite integral, e.g. work (W), heat (Q) atomisation energy enthalpy of formation dissociation energy

[A 1 / [A I,

[BlI[Bl,

[DI / P I ,

Page 11: Advanced Thermodynamics for Engineers

Symbols

fraction of reaction potential difference, voltage Seebeck coefficient for material pair A, B eddy diffusivity ratio of specific heats (c,/c,) electrical conductivity dynamic viscosity Joule-Thomson coefficient chemical potential electrochemical potential kinematic viscosity Peltier coefficient for material pair A, B entropy generation per unit volume Thomson coefficient stoichiometric Coefficient

efficiency equivalence ratio density specific exergy inner electric potential of a phase temperature ratio exergy

[CI/[Cle

Suffices

0 actual av b b B

C C [ l e f f fg

C

g g h H i i in isen

dead state conditions value from actual cycle, as opposed to ideal cycle available (as in energy) backward (reaction) burned (products of combustion) boiler at critical point, e.g. pressure, temperature, specific volume cold (as in temperature of reservoir) compressor equilibrium molar density forward (reaction) value for saturated liquid, e.g. h, = enthalpy of saturated liquid difference between properties on saturated vapour and saturated liquid lines, i.e.

value for saturated gas, e.g. h, = enthalpy of saturated gas gaseous state (as in reactants or products) constant enthalpy hot (as in temperature of reservoir) inversion ith constituent into system isentropic (as in a process)

h,= h,- hf

Page 12: Advanced Thermodynamics for Engineers

Symbols xix

1 1 1 max mol net 0 oc out P P P rejected

R R R R

E S

S

S SUIT

SYS T T th U

U unav univ use V

latent energy required for evaporation (equals h, or ufg) liquid state (as in reactants or products) laminZiI maximum value, or maximum useful work specific property based on mols net (as in work output from system) overall open circuit out of system at constant pressure products Pump not used by cycle (usually energy) energy contained in molecules by resonance rational (as in efficiency) reactants reduced properties (in Law of Corresponding States) reversible shaft (as in work) at constant entropy surroundings system turbine at constant temperature thermal useful (as in work) unburned (as of reactants) unavailable (as in energy) universe (Le. system + surroundings) useful (as in work) at constant specific volume

Page 13: Advanced Thermodynamics for Engineers

Contents

Preface structure Symbols

1 State of Equilibrium

1.1 Equilibrium of a thermodynamic system 1.2 Helmholtz energy (Helmholtz function) 1.3 Gibbs energy (Gibbs function) 1.4 1.5 Concluding remarks Problems

The use and significance of the Helmholtz and Gibbs energies

2 Availability and Exergy

2.1 Displacement work 2.2 Availability 2.3 Examples 2.4 Available and non-available energy 2.5 Irreversibility 2.6 2.7 2.8 2.9 Exergy 2.10 2.1 1 Concluding remarks Problems

Graphical representation of available energy and irreversibility Availability balance for a closed system Availability balance for an open system

The variation of flow exergy for a perfect gas

3 Pinch Technology

3.1 3.2 3.3 Concluding remarks Problems

A heat transfer network without a pinch problem A heat transfer network with a pinch point

ix ... X l l l

xv

1

2 5 6 6 9

10

13

13 14 15 21 21 25 27 34 36 42 43 43

47

49 56 60 61

Page 14: Advanced Thermodynamics for Engineers

vi Contents

4 Rational Efficiency of a Powerplant

4.1 4.2 Rational efficiency 4.3 Rankinecycle 4.4 Examples 4.5 Concluding remarks Problems

The influence of fuel properties on thermal efficiency

64

64 65 69 71 82 82

5 Efficiency of Heat Engines at Maximum Power

5.1

5.2

5.3 Concluding remarks Problems

Efficiency of an internally reversible heat engine when producing maximum power output Efficiency of combined cycle internally reversible heat engines when producing maximum power output

6 General Thermodynamic Relationships (single component systems, or systems of constant composition 1 6.1 The Maxwell relationships 6.2 6.3 Tdr relationships 6.4 6.5 The Clausius-Clapeyron equation 6.6 Concluding remarks Problems

Uses of the thermodynamic relationships

Relationships between specific heat capacities

7 Equations of State 7.1 Ideal gas law 7.2 7.3 Law of corresponding states 7.4 7.5 Concluding remarks Problems

Van der Waals’ equation of state

Isotherms or isobars in the two-phase region

8 Liquefaction of Gases

8.1 8.2 8.3 The Joule-Thomson effect 8.4 Linde liquefaction plant 8.5 8.6 Concluding remarks Problems

Liquefaction by cooling - method (i) Liquefaction by expansion - method (ii)

Inversion point on p-v-T surface for water

9 Thermodynamic Properties of Ideal Gases and Ideal Gas Mixtures of Constant Composition 9.1 Molecular weights

85

85

92 96 96

100 100 104 108 111 115 118 118

121

121 123 125 129 131 132

135

135 140 141 148 150 155 155

158 158

Page 15: Advanced Thermodynamics for Engineers

9.2 9.3 9.4 Mixtures of ideal gases 9.5 Entropy of mixtures 9.6 Concluding remarks Problems

State equation for ideal gases Tables of u(T) and h(T) against T

10 Thermodynamics of Combustion

10.1 Simple chemistry 10.2 10.3 10.4

10.5 Combustion processes 10.6 Examples 10.7 Concluding remarks Problems

Combustion of simple hydrocarhn fuels Heats of formation and heats of reaction Application of the energy equation to the combustion process - a macroscopic approach

11 Chemistry of Combustion

1 1.1 1 1.2 Energy of formation 1 1.3 Enthalpy of reaction 1 1.4 Concluding remarks

Bond energies and heats of formation

12 Chemical Equilibrium and Dissociation

12.1 Gibbs energy 12.2 Chemical potential, p 12.3 Stoichiometry 12.4 Dissociation 12.5 12.6 12.7 12.8 12.9 12.10 Dissociation calculations for the evaluation of nitric oxide 12.11 Dissociation problems with two, or more, degrees of dissociation 12.12 Concluding remarks Problems

Calculation of chemical equilibrium and the law of mass action Variation of Gibbs energy with composition Examples of the significance of Kp The Van? Hoff relationship between equilibrium constant and heat of reaction The effect of pressure and temperature on degree of dissociation

13 The Effect of Dissociation on Combustion Parameters

13.1 Calculation of combustion both with and without dissociation 13.2 The basic reactions 13.3 The effect of dissociation on peak pressure 13.4 The effect of dissociation on peak temperature 13.5 The effect of dissociation on the composition of the products 13.6 The effect of fuel on composition of the products 13.7 The formation of oxides of nitrogen

159 164 172 175 178 178

182

184 185 187

188 192 195 205 205

208

208 210 216 216

218

218 220 221 222 225 229 23 1 238 239 242 245 259 259

265

267 267 268 268 269 272 273

Page 16: Advanced Thermodynamics for Engineers

viii Contents

14 Chemical Kinetics

14.1 Introduction 14.2 Reaction rates 14.3 14.4 Chemical kinetics of NO 14.5 14.6 14.7 Concluding remarks Problems

Rate constant for reaction, k

The effect of pollutants formed through chemical kinetics Other methods of producing power from hydrocarbon fuels

276

276 276 279 280 286 288 289 289

15 Combustion and Flames

15.1 Introduction 15.2 Thermodynamics of combustion 15.3 Explosion limits 15.4 Flames 15.5 Flammability limits 15.6 Ignition 15.7 Diffusion flames 15.8 Engine combustion systems 15.9 Concluding remarks Problems

16 Irreversible Thermodynamics 16.1 Introduction 16.2 Definition of irreversible or steady state thermodynamics 16.3 Entropy flow and entropy production 16.4 Thermodynamic forces and thermodynamic velocities 16.5 Onsager’s reciprocal relation 16.6 The calculation of entropy production or entropy flow 16.7 Thermoelectricity - the application of irreversible thermodynamics to a

thermocouple 16.8 Diffusion and heat transfer 16.9 Concluding remarks Problems

17 Fuel Cells 17.1 Electric cells 17.2 Fuel cells 17.3 17.4 17.5 Concluding remarks Problems

Efficiency of a fuel cell Thermodynamics of cells working in steady state

29 1

29 1 292 294 296 303 304 305 307 314 314

316 316 317 317 318 319 321

322 332 342 342

345 346 35 1 358 359 361 361

Bibliography 363

Index (including Index of tables of properties) 369

Page 17: Advanced Thermodynamics for Engineers

State of Equilibrium

Most texts on thermodynamics restrict themselves to dealing exclusively with equilibrium thermodynamics. This book will also focus on equilibrium thermodynamics but the effects of making this assumption will be explicitly borne in mind. The majority of processes met by engineers are in thermodynamic equilibrium, but some important processes have to be considered by non-equilibrium thermodynamics. Most of the combustion processes that generate atmospheric pollution include non-equilibrium effects, and carbon monoxide (CO) and oxides of nitrogen (NO,.) are both the result of the inability of the system to reach thermodynamic equilibrium in the time available.

There are four kinds of equilibrium, and these are most easily understood by reference to simple mechanical systems (see Fig 1.1).

(i) Stable equilibrium w Marble in bowl. For stable equilibrium AS),* < 0 and A E ) s > 0. (AS is the sum of Taylor’s series terms). Any deflection causes motion back towards equilibrium position. * Discussed later.

( i i ) Neutral equilibrium

Marble in trough. AS), = 0 and A E ) s = 0 along trough axis. Marble in equilibrium at any position in x-direction.

(iii) Unstable equilibrium Marble sitting on maximum point of surface. A S ) E > 0 and A E ) s < 0. Any movement causes further motion from ‘equilibrium’ position.

Fig. 1.1 States of equilibrium -

Page 18: Advanced Thermodynamics for Engineers

2 State of equilibrium

(iv) Metastable equilibrium

Marble in higher of two troughs. Infinitesimal variations of position cause return to equilibrium - larger variations cause movement to lower level.

Fig. 1.1 Continued

*The difference between AS and dS Consider Taylor’s theorem

Thus dS is the first term of the Taylor’s series only. Consider a circular bowl at the position where the tangent is horizontal. Then

1 d2S d2S However AS = dS + - - Ax2 + ... # 0, because - etc are not zero.

2 dr2 dr2

Hence the following statements can be derived for certain classes of problem

stable equilibrium (ds), = 0

(As) E < 0 neutral equilibrium (dS)E = 0

(AS), = 0 unstable equilibrium (dS), = 0

( A s ) E > 0

(see Hatsopoulos and Keenan, 1972).

1.1 The type of equilibrium in a mechanical system can be judged by considering the variation in energy due to an infinitesimal disturbance. If the energy (potential energy) increases

Equilibrium of a thermodynamic system

Page 19: Advanced Thermodynamics for Engineers

Equilibrium of a thermodynamic system 3

F = T l Fig. 1.2 Heat transfer between two blocks

then the system will return to its previous state, if it decreases it will not return to that state.

A similar method for examining the equilibrium of thermodynamic systems is required. This will be developed from the Second Law of Thermodynamics and the definition of entropy. Consider a system comprising two identical blocks of metal at different temperatures (see Fig 1.2), but connected by a conducting medium. From experience the block at the higher temperature will transfer ‘heat’ to that at the lower temperature. If the two blocks together constitute an isolated system the energy transfers will not affect the total energy in the system. If the high temperature block is at an temperature TI and the other at T2 and if the quantity of energy transferred is SQ then the change in entropy of the high temperature block is

SQ TI

1 -

and that of the lower temperature block is

SQ a,=+- T2

Both eqns (1.1) and (1.2) contain the assumption that the heat transfers from block 1, and into block 2 are reversible. If the transfers were irreversible then eqn (1.1) would become

SQ ds, > -- TI

and eqn (1.2) would be

SQ ds, > +- 7-2

(l.la)

(1.2a)

Since the system is isolated the energy transfer to the surroundings is zero, and hence the change of entropy of the surroundings is zero. Hence the change in entropy of the system is equal to the change in entropy of the universe and is, using eqns (1.1) and (1.2),

Since TI > T,, then the change of entropy of both the system and the universe is dS = (6Q/T2Tl)(T, - T2) > 0. The same solution, namely dS > 0, is obtained from eqns (l.la) and (1.2a).

The previous way of considering the equilibrium condition shows how systems will tend to go towards such a state. A slightly different approach, which is more analogous to the

Page 20: Advanced Thermodynamics for Engineers

4 State of equilibrium

one used to investigate the equilibrium of mechanical systems, is to consider these two blocks of metal to be in equilibrium and for heat transfer to occur spontaneously (and reversibly) between one and the other. Assume the temperature change in each block is 6T, with one temperature increasing and the other decreasing, and the heat transfer is SQ. Then the change of entropy, CIS, is given by

- 6Q ( T - S T - T - 6 T ) &=---- SQ SQ T + 6T T - 6T ( T + 6 T ) ( T - 6T)

dT (-26T) -26Q - - SQ -

T 2 + 6T2 T 2 (1.4)

This means that the entropy of the system would have decreased. Hence maximum entropy is obtained when the two blocks are in equilibrium and are at the same temperature. The general criterion of equilibrium according to Keenan (1963) is as follows.

For stability of any system it is necessary and sufficient that, in all possible variations of the state of the system which do not alter its energy, the variation of entropy shall be negative.

This can be stated mathematically as

AS), < 0 (1.5)

It can be seen that the statements of equilibrium based on energy and entropy, namely A E ) , > 0 and AS), < 0, are equivalent by applying the following simple analysis. Consider the marble at the base of the bowl, as shown in Fig l.l(i): if it is lifted up the bowl, its potential energy will be increased. When it is released it will oscillate in the base of the bowl until it comes to rest as a result of ‘friction’, and if that ‘friction’ is used solely to raise the temperature of the marble then its temperature will be higher after the process than at the beginning. A way to ensure the end conditions, i.e. the initial and final conditions, are identical would be to cool the marble by an amount equivalent to the increase in potential energy before releasing it. This cooling is equivalent to lowering the entropy of the marble by an amount AS, and since the cooling has been undertaken to bring the energy level back to the original value this proves that AE), > 0 and AS), < 0.

Equilibrium can be defined by the following statements:

(i)

(ii)

(iii)

if the properties of an isolated system change spontaneously there is an increase in the entropy of the system; when the entropy of an isolated system is at a maximum the system is in equilibrium; if, for all the possible variations in state of the isolated system, there is a negative change in entropy then the system is in stable equilibrium.

These conditions may be written mathematically as:

(i) dS), > 0 spontaneous change (unstable equilibrium) (ii) dS), = 0 equilibrium (neutral equilibrium) (iii) AS), < 0 criterion of stability (stable equilibrium)

Page 21: Advanced Thermodynamics for Engineers

Helmholtz energy (Helmholtz function) 5

1.2 Helmholtz energy (Helmholtz function) There are a number of ways of obtaining an expression for Helmholtz energy, but the one based on the Clausius derivation of entropy gives the most insight.

In the previous section, the criteria for equilibrium were discussed and these were derived in terms of A S ) E . The variation of entropy is not always easy to visualise, and it would be more useful if the criteria could be derived in a more tangible form related to other properties of the system under consideration. Consider the arrangements in Figs 1.3(a) and (b). Figure 1.3(a) shows a System A, which is a general system of constant composition in which the work output, 6W, can be either shaft or displacement work, or a combination of both. Figure 1.3(b) is a more specific example in which the work output is displacement work, p 6V; the system in Fig 1.3(b) is easier to understand.

------------. ,-----

"~, System B ,I'

SvstemA

Reservoir To (b)

Fig. 1.3 Maximum work achievable from a system

In both arrangements, System A is a closed system (i.e. there are no mass transfers), which delivers an infinitesimal quantity of heat SQ in a reversible manner to the heat engine E,. The heat engine then rejects a quantity of heat SQ, to a reservoir, e.g. the atmosphere, at temperature To.

Let dE, dV and dS denote the changes in internal energy, volume and entropy of the system, which is of constant, invariant composition. For a specified change of state these quantities, which are changes in properties, would be independent of the process or work done. Applying the First Law of Thermodynamics to System A gives

6W= -dE + S Q (1.6)

If the heat engine (E,) and System A are considered to constitute another system, System B, then, applying the First Law of Thermodynamics to System B gives

(1.7)

where 6 W + 6 W, = net work done by the heat engine and System A. Since the heat engine is internally reversible, and the entropy flow on either side is equal, then

6 W,,, = 6 W + 6 W, = - dE + 6 Q,

Page 22: Advanced Thermodynamics for Engineers

6 State of equilibrium

and the change in entropy of System A during this process, because it is reversible, is dS = SQ,/T. Hence

because To = constant SW,, = -dE + To dS

= -d(E - TOS) (1.9)

The expression E-ToS is called the Helmholtz energy or Helmholtz function. In the absence of motion and gravitational effects the energy, E, may be replaced by the intrinsic internal energy, U , giving

SW,,= -d(U - TOS) (1.10) The significance of SW,, will now be examined. The changes executed were considered to be reversible and SW,,, was the net work obtained from System B (i.e. System A +heat engine ER). Thus, SW,,, must be the maximum quantity of work that can be obtained from the combined system. The expression for SW is called the change in the Helmholtz energy, where the Helmholtz energy is defined as

F = U - T S (1.11)

Helmholtz energy is a property which has the units of energy, and indicates the maximum work that can be obtained from a system. It can be seen that this is less than the internal energy, U, and it will be shown that the product TS is a measure of the unavailable energy.

1.3 Gibbs energy (Gibbs function) In the previous section the maximum work that can be obtained from System B, comprising System A and heat engine ER, was derived. It was also stipulated that System A could change its volume by SV, and while it is doing this it must perform work on the atmosphere equivalent to po SV, where po is the pressure of the atmosphere. This work detracts from the work previously calculated and gives the maximum useful work, SW,, as

SW, = SW,, - P O dV (1.12)

if the system is in pressure equilibrium with surroundings.

SW,= -d(E- ToS)-p, dV = -d(E +poV- TOS)

because p o = constant. Hence SW, = -d(H - 7's) (1.13)

The quantity H - TS is called the Gibbs energy, Gibbs potential, or the Gibbs function, G. Hence

G = H - T S (1.14)

Gibbs energy is a property which has the units of energy, and indicates the maximum useful work that can be obtained from a system. It can be seen that this is less than the enthalpy, H , and it will be shown that the product TS is a measure of the unavailable energy.

1.4 The use and significance of the Helmholtz and Gibbs energies It should be noted that the definitions of Helmholtz and Gibbs energies, eqns (1.11) and (1.14), have been obtained for systems of invariant composition. The more general form of

Page 23: Advanced Thermodynamics for Engineers

The use and significance of the Helmholtz and Gibbs energies 7

these basic thermodynamic relationships, in differential form, is

dU = T dS -p dV+ C pi dn, d H = T dS+ V dp + C mi dn, d F = -S dT-p d V + C ,LA, dn, d G = - S d T + V d p + C p , d n , (1.15)

The additional term, pi dn,, is the product of the chemical potential of component i and the change of the amount of substance (measured in moles) of component i. Obviously, if the amount of substance of the constituents does not change then this term is zero. However, if there is a reaction between the components of a mixture then this term will be non-zero and must be taken into account. Chemical potential is introduced in Chapter 12 when dissociation is discussed; it is used extensively in the later chapters where it can be seen to be the driving force of chemical reactions.

1.4.1 HELMHOLTZ ENERGY

(i) The change in Helmholtz energy is the maximum work that can be obtained from a closed system undergoing a reversible process whilst remaining in temperature equilibrium with its surroundings.

(ii) A decrease in Helmholtz energy corresponds to an increase in entropy, hence the minimum value of the function signifies the equilibrium condition.

(iii) A decrease in entropy corresponds to an increase in F; hence the criterion dF), > 0 is that for stability. This criterion corresponds to work being done on the system. For a constant volume system in which W = 0, d F = 0. For reversible processes, F, = F,; for all other processes there is a decrease in Helmholtz energy. The minimum value of Helmholtz energy corresponds to the equilibrium condition.

(iv) (v)

(vi)

1.4.2 GIBBS ENERGY

(i) The change in Gibbs energy is the maximum useful work that can be obtained from a system undergoing a reversible process whilst remaining in pressure and temperature equilibrium with its surroundings.

(ii) The equilibrium condition for the constraints of constant pressure and temperature can be defined as:

(1) dG),,, < 0 spontaneous change

(2) dG),,T = 0 equilibrium

(3) AG&, > 0 criterion of stability.

(iii) The minimum value of Gibbs energy corresponds to the equilibrium condition.

Page 24: Advanced Thermodynamics for Engineers

8 State of equilibrium

1.4.3 THERMODYNAMICS

EXAMPLES OF DIFFERENT FORMS OF EQUILJBRIUM MET IN

Stable equilibrium is the most frequently met state in thermodynamics, and most systems exist in this state. Most of the theories of thermodynamics are based on stable equilibrium, which might be more correctly named thermostatics. The measurement of thermodynamic properties relies on the measuring device being in equilibrium with the system. For example, a thermometer must be in thermal equilibrium with a system if it is to measure its temperature, which explains why it is not possible to assess the temperature of something by touch because there is heat transfer either to or from the fingers - the body ‘measures’ the heat transfer rate. A system is in a stable state if it will permanently stay in this state without a tendency to change. Examples of this are a mixture of water and water vapour at constant pressure and temperature; the mixture of gases from an internal combustion engine when they exit the exhaust pipe; and many forms of crystalline structures in metals. Basically, stable equilibrium states are defined by state diagrams, e.g. the p-v-T diagram for water, where points of stable equilibrium are defined by points on the surface; any other points in the p-v-T space are either in unstable or metastable equilibrium. The equilibrium of mixtures of elements and compounds is defined by the state of maximum entropy or minimum Gibbs or Helmholtz energy; this is discussed in Chapter 12. The concepts of stable equilibrium can also be used to analyse the operation of fuel cells and these are considered in Chapter 17.

Another form of equilibrium met in thermodynamics is metastable equilibrium. This is where a system exists in a ‘stable’ state without any tendency to change until it is perturbed by an external influence. A good example of this is met in combustion in spark-ignition engines, where the reactants (air and fuel) are induced into the engine in a pre-mixed form. They are ignited by a small spark and convert rapidly into products, releasing many thousands of times the energy of the spark used to initiate the combustion process. Another example of metastable equilibrium is encountered in the Wilson ‘cloud chamber’ used to show the tracks of a particles in atomic physics. The Wilson cloud chamber consists of super-saturated water vapour which has been cooled down below the dew-point without condensation - it is in a metastable state. If an a particle is introduced into the chamber it provides sufficient perturbation to bring about condensation along its path. Other examples includz explosive boiling, which can occur if there are not sufficient nucleation sites to induce sufficient bubbles at boiling point to induce normal boiling, and some of the crystalline states encountered in metallic structures.

Unstable states cannot be sustained in thermodynamics because the molecular movement will tend to perturb the systems and cause them to move towards a stable state. Hence, unstable states are only transitory states met in systems which are moving towards equilibrium. The gases in a combustion chamber are often in unstable equilibrium because they cannot react quickly enough to maintain the equilibrium state, which is defined by minimum Gibbs or Helmholtz energy. The ‘distance’ of the unstable state from the state of stable equilibrium defines the rate at which the reaction occurs; this is referred to as rate kinetics, and will be discussed in Chapter 14. Another example of unstable ‘equilibrium’ occurs when a partition is removed between two gases which are initially separated. These gases then mix due to diffusion, and this mixing is driven by the difference in chemical potential between the gases; chemical potential is introduced in Chapter 12 and the process

Page 25: Advanced Thermodynamics for Engineers

Concluding remarks 9

of mixing is discussed in Chapter 16. Some thermodynamic situations never achieve stable equilibrium, they exist in a steady state with energy passing between systems in stable equilibrium, and such a situation can be analysed using the techniques of irreversible thermodynamics developed in Chapter 16.

1.4.4 PRESSURE AND TEMPERATURE

SIGNIFICANCE OF THE MINIMUM GIBBS ENERGY AT CONSTANT

It is difficult for many engineers readily to see the significance of Gibbs and Helmholtz energies. If systems are judged to undergo change while remaining in temperature and pressure equilibrium with their surroundings then most mechanical engineers would feel that no change could have taken place in the system. However, consideration of eqns (1.15) shows that, if the system were a multi-component mixture, it would be possible for changes in Gibbs (or Helmholtz) energies to take place if there were changes in composition. For example, an equilibrium mixture of carbon dioxide, carbon monoxide and oxygen could change its composition by the carbon dioxide brealung down into carbon monoxide and oxygen, in their stoichiometric proportions; this breakdown would change the composition of the mixture. If the process happened at constant temperature and pressure, in equilibrium with the surroundings, then an increase in the Gibbs energy, G , would have occurred; such a process would be depicted by Fig 1.4. This is directly analogous to the marble in the dish, which was discussed in the introductory remarks to this section.

% Carbon dioxide in mixture

Fig. 1.4 Variation of Gibbs energy with chemical composition, for a system in temperature and pressure equilibrium with the environment

1.5 Concluding remarks

This chapter has considered the state of equilibrium for thermodynamic systems. Most systems are in equilibrium, although non-equilibrium situations will be introduced in Chapters 14, 16 and 17.

Page 26: Advanced Thermodynamics for Engineers

10 State of equilibrium

It has been shown that the change of entropy can be used to assess whether a system is in a stable state. Two new properties, Gibbs and Helmholtz energies, have been introduced and these can be used to define equilibrium states. These energies also define the maximum amount of work that can be obtained from a system.

PROBLEMS 1 Determine the criteria for equilibrium for a thermally isolated system at (a) constant

volume; (b) constant pressure. Assume that the system is: (i) constant, and invariant, in composition; (ii) variable in composition.

2 Determine the criteria for isothermal equilibrium of a system at (a) constant volume, and (b) constant pressure. Assume that the system is:

(i) constant, and invariant, in composition; (ii) variable in composition.

3 A system at constant pressure consists of 10 kg of air at a temperature of lo00 K. This is connected to a large reservoir which is maintained at a temperature of 300 K by a reversible heat engine. Calculate the maximum amount of work that can be obtained from the system. Take the specific heat at constant pressure of air, cp, as 0.98 H/kg K.

[3320.3 kJ]

4 A thermally isolated system at constant pressure consists of 10 kg of air at a temperature of 1000 K and 10 kg of water at 300 K, connected together by a heat engine. What would be the equilibrium temperature of the system if (a) the heat engine has a thermal efficiency of zero; (b) the heat engine is reversible?

{Hint: consider the definition of equilibrium defined by the entropy change of the system. } Assume

for water: c, = 4.2 kJ/kg K

C, = 0.7 kJ/kg K K = cp/c, = 1.0

for air: K = cP/c, = 1.4

[432.4 K; 376.7 K]

5 A thermally isolated system at constant pressure consists of 10 kg of air at a temperature of 1000 K and 10 kg of water at 300 K, connected together by a heat engine. What would be the equilibrium temperature of the system if the maximum thermal efficiency of the engine is only 50%?

Assume for water: c, = 4.2 kJ /kg K

C, = 0.7 kJ/kg K 7~ = c ~ / c , = 1.4

K = c,/c, = 1.0 for air:

r385.1 K]

Page 27: Advanced Thermodynamics for Engineers

Problems 11

6 Show that if a liquid is in equilibrium with its own vapour and an inert gas in a closed vessel, then

dPv Pv

dP P - = -

where p, is the partial pressure of the vapour, p is the total pressure, p, is the density of the vapour and p1 is the density of the liquid.

7 An incompressible liquid of specific volume v,, is in equilibrium with its own vapour and an inert gas in a closed vessel. The vapour obeys the law

P ( V - b)=%T

Show that

- = - I(P - P0)Y - (Pv -Po)b) (L) i T

where po is the vapour pressure when no inert gas is present, and p is the total pressure.

8 (a) Describe the meaning of the term thermodynamic equilibrium. Explain how entropy can be used as a measure of equilibrium and also how other properties can be developed that can be used to assess the equilibrium of a system.

If two phases of a component coexist in equilibrium (e.g. liquid and vapour phase H,O) show that

T - = - dP 1

dT Vfg

where

T = temperature, p = pressure, 1 = latent heat, vfg = difference between liquid and vapour phases. and

Show the significance of this on a phase diagram. (b) The melting point of tin at a pressure of 1 bar is 505 K, but increases to 508.4 K at 1000 bar. Evaluate

(i) the change of density between these pressures, and (ii) the change in entropy during melting.

The latent heat of fusion of tin is 58.6 kJ/kg. [254 100 kg/m3; 0.1 157 kJ/kg K]

9 Show that when different phases are in equilibrium the specific Gibbs energy of each phase is equal.

Using the following data, show the pressure at which graphite and diamond are in

Page 28: Advanced Thermodynamics for Engineers

12 State of equilibrium

equilibrium at a temperature of 25°C. The data for these two phases of carbon at 25°C and 1 bar are given in the following table:

Graphite Diamond

Specific Gibbs energy, g/(kJ/kg) 0 269 Specific volume, v/(m3/kg) 0.446 x 1 0 - ~ 0.285 x Isothermal compressibility, k/(bar-') 2.96 x 0.158 x

It may be assumed that the variation of kv with pressure is negligible, and the lower

[ 17990 bar] value of the solution may be used.

10 Van der Waals' equation for water is given by

0.004619T 0.017034 - P = v - 0.0016891 V 2

where p = pressure (bar), v = specific volume (m3/kmol), T = temperature (K).

Draw a p - v diagram for the following isotherms: 250"C, 270"C, 300°C, 330°C, 374"C, 390°C.

Compare the computed specific volumes with Steam Table values and explain the differences in terms of the value of pcvc/%Tc.

Page 29: Advanced Thermodynamics for Engineers

Availability and Exergy

Many of the analyses performed by engineers are based on the First Law of Thermo- dynamics, which is a law of energy conservation. Most mechanical engineers use the Second Law of Thermodynamics simply through its derived property - entropy ( S ) . However, it is possible to introduce other ‘Second Law’ properties to define the maximum amounts of work achievable from certain systems. Previously, the properties Helmholtz energy ( F ) and Gibbs energy ( G ) were derived as a means of assessing the equilibrium of various systems. This section considers how the maximum amount of work available from a system, when interacting with surroundings, can be estimated. This shows, as expected, that all the energy in a system cannot be converted to work: the Second Law stated that it is impossible to construct a heat engine that does not reject energy to the surroundings.

2.1 Displacement work The work done by a system can be considered to be made up of two parts: that done against a resisting force and that done against the environment. This can be seen in Fig 2.1. The pressure inside the system, p , is resisted by a force, F , and the pressure of the environment. Hence, for System A, which is in equilibrium with the surroundings

p A = F + p d (2.1) where A is the area of cross-section of the piston.

Svstem A

Fig. 2.1 Forces acting on a piston

If the piston moves a distance dx, then the work done by the various components shown in Fig 2.1 is

pA d x = F d x + p J dx (2.2) where PA dx = p dV = SW,,, = work done by the fluid in the system, F dx = SW,, = work done against the resisting force, and pJ dx = po dV = SW,, = work done against the surroundings.

Page 30: Advanced Thermodynamics for Engineers

14 Availability and exergy

Hence the work done by the system is not all converted into useful work, but some of it is used to do displacement work against the surroundings, i.e.

sw,,, = SW, + SW,, (2.3)

which can be rearranged to give

sw, = SW,,, - SW,, (2.4)

2.2 Availability

It was shown above that not all the displacement work done by a system is available to do useful work. This concept will now be generalised to consider all the possible work outputs from a system that is not in thermodynamic and mechanical equilibrium with its surroundings (i.e. not at the ambient, or dead state, conditions).

Consider the system introduced earlier to define Helmholtz and Gibbs energy: this is basically the method that was used to prove the Clausius inequality.

Figure 2.2(a) shows the general case where the work can be either displacement or shaft work, while Fig 2.2(b) shows a specific case where the work output of System A is displacement work. It is easier to follow the derivation using the specific case, but a more general result is obtained from the arrangement shown in Fig 2.2(a).

System B

6W

6WR *

Reservoir To (4

.......... ,.--

'\, System B Svstem A ',,

0) e7 Reservoir To

Fig. 2.2 System transferring heat to a reservoir through a reversible heat engine

First consider that System A is a constant volume system which transfers heat with the

(2.5)

Let System B be System A plus the heat engine E,. Then applying the First Law to

surroundings via a small reversible heat engine. Applying the First Law to the System A

d U = SQ- SW= SQ- SWS

where SW, indicates the shaft work done (e.g. System A could contain a turbine).

system B gives

d U = C (SQ-SW)=SQ,-(SWs+SWR) (2.6)

Page 31: Advanced Thermodynamics for Engineers

Examples 15

As System A transfers energy with the surroundings it under goes a change of entropy defined by

because the heat engine transferring the heat to the surroundings is reversible, and there is no change of entropy across it. Hence

(SW~+dWR)=-(dU-TodS) (2.8)

As stated previously, (6Ws + 6WR) is the maximum work that can be obtained from a constant volume, closed system when interacting with the surroundings. If the volume of the system was allowed to change, as would have to happen in the case depicted in Fig 2.2(b), then the work done against the surroundings would be po dV, where po is the pressure of the surroundings. This work, done against the surroundings, reduces the maximum useful work from the system in which a change of volume takes place to 6W + SW, - p o dV, where 6W is the sum of the shaft work and the displacement work.

Hence, the maximum useful work which can be achieved from a closed system is

6 W + SW, = - (dU +Po dV - To dS) (2.9)

This work is given the symbol dA. Since the surroundings are at fixed pressure and temperature (Le. po and To are constant) dA can be integrated to give

A = U + p o V - TOS (2.10)

A is called the non-flow availability function. Although it is a combination of properties, A is not itself a property because it is defined in relation to the arbitrary datum values of p o and To. Hence it is not possible to tabulate values of A without defining both these datum levels. The datum levels are what differentiates A from Gibbs energy G. Hence the maximum useful work achievable from a system changing state from 1 to 2 is given by

W,,= -AA= -(A2-Al)=Al-A2 (2.11)

The specific availability, a , i.e. the availability per unit mass is

a = u +pov - Tos (2.12a)

If the value of a were based on unit amount of substance (Le. kmol) it would be referred to as the molar availability.

The change of specific (or molar) availability is

h a = a2 - a , = (u2 + pov2 - Tos2) - ( u l + pov, - Tosl)

= (h2 + v2( Po - P d ) - ( h l + V I ( Po - P I ) ) - To(% - $1) (2.12b)

2.3 Examples

Example 1 : reversible work from a piston cylinder arrangement (this example is based on Haywood (1980))

System A, in Fig 2.1, contains air at a pressure and temperature of 2 bar and 550 K respectively. The pressure is maintained by a force, F , acting on the piston. The system is taken from state 1 to state 2 by the reversible processes depicted in Fig 2.3, and state 2 is

Page 32: Advanced Thermodynamics for Engineers

16 Availability and exergy

equal to the dead state conditions with a pressure, p,,, and temperature, To, of 1 bar and 300 K respectively. Evaluate the following work terms assuming that the air is a perfect gas and that cp = 1.005 W/kg K and the ratio of specific heats K = 1.4.

(a) The air follows the process 1-a-2 in Fig 2.3, and transfers heat reversibly with the environment during an isobaric process from l-a. Calculate the following specific work outputs for processes 1-a and a-2:

(i) (ii) (iii) (iv)

the work done by the system, 6 W,,,; the work done against the surroundings, SW,,; the useful work done against the resisting force F , SW,; the work done by a reversible heat engine operating between the system and the surroundings, 6 WR .

Then evaluate for the total process, 1-2, the following parameters:

(v) the gross work done by the system, C(SW,,, + 6WR); (vi) the net useful work output against the environment, C(SW, + 6WR); (vii) the total displacement work against the environment, C( SW,,); (viii) the work term po(v2 - v,).

2

Entropy, S

Fig. 2.3 Processes undergone by the system

Solution

It is necessary to evaluate the specific volume at the initial condition, state 1.

p , v , = RT,, hence v, = R T , / p ,

The gas constant, R , is given by R = ((K - l)/x)cp = 0.4 x 1.005/1.4 = 0.287 kJ/kgK, and hence

0.287 x lo3 x 550 v, = = 0.7896 m3/kg

2 x io5

Page 33: Advanced Thermodynamics for Engineers

Examples 17

The intermediate temperature, T,, should now be evaluated. Since the process a-2 is isentropic then Ta must be isentropically related to the final temperature, T,. Hence

( p )-m (:0.4/1.4

T = T a 2 2 = 3 0 0 - = 365.7 K P2

365.7 v1 = - V I = 0.6649~1 PlTa Thus v a = -

PaTl 550 Consider process 1 -a:

wsYs - a = IIap dv = pa(va - vl), for an isobaric process

2 io5

io3 (0.6649 - 1 .O) x 0.7896 = -52.96 kJ/kg -- -

1 io5

io3 Ws,ll-a = IIaPodv = ~ (0.6649 - 1.0) x 0.7896

= -26.46kJ/kg The useful work done by the system against the resisting force, F , is the difference between the two work terms given above, and is

wuse 1 1 - a = wsys 1 1 - a - ws, I l - a = -26.46 kJ/kg The work terms derived above relate to the mechanical work that can be obtained from

the system as it goes from state 1 to state a. However, in addition to this mechanical work the system could also do thermodynamic work by transfemng energy to the surroundings through a reversible heat engine. This work can be evaluated in the following way. In going from state 1 to state a the system has had to transfer energy to the surroundings because the temperature of the fluid has decreased from 550 K to 365.7 K. This heat transfer could have been simply to a reservoir at a temperature below T,, in which case no useful work output would have been achieved. It could also have been to the environment through a reversible heat engine, as shown in Fig 2.2(b). In the latter case, useful work would have been obtained, and this would be equal to

SwR = -vRSQ, where vR is the thermal efficiency of a reversible heat engine operating between T and To, and SQ is the heat transfer to the system. Thus

Hence, the work output obtainable from this reversible heat engine as the system changes state from 1 to a is

a T - T o - SQ = -1, - c, dT

T WR = -

( T , - T , ) - To In -

365.7

500 (365.7 - 550) - 300 In ___

Page 34: Advanced Thermodynamics for Engineers

18 Availability and exergy

Now consider process a-2. First, evaluate the specific volume at 2, v2:

p lT2 2 x 300 ~2T1 1 x 550

v2 = - V , ___ v1 = 1 .0909~~

The expansion from a to 2 is isentropic and hence obeys the law pv" = constant. Thus the system work is

(1 x 1.0909 - 2 x 0.6649) 0.7896 x lo5 X = 47.16 kJ/kg

io3 wsysla-2 =

1 - 1.4

2 1 io5 ~ s m r 1 a - 2 = j ~ ~ o d v = p (1.0909 - 0.6649) x 0.7896 = 33.63 W/kg

io3

wuseIa-2 = wsysIa-2-~sumIa-2= 13*52kJ/kg

The energy available to drive a reversible heat engine is zero for this process because it is adiabatic; hence

wR l a - 2 =

The work terms for the total process from 1 to 2 can be calculated by adding the terms for the two sub-processes. This enables the solutions to questions (v) to (viii) to be obtained.

The gross work done by the system is

w ~ ~ ~ = ~ ( w ~ ~ ~ + wR)= -52.96+47.16+62.18+0=56.38 W/kg

The net useful work done by the system is

W,, ,~=E(W,+ wR)= -26.46+ 13.52+62.18=49.24 kJ/kg

The total displacement work done against the surroundings is

w,,= -26.46+33.63=7.17 W/kg

The displacement work of the surroundings evaluated from

p0(v2 - vl ) = 1 x lo5 x (1.0909 - 1) x 0.7896/103 = 7.17 kJ/kg

Example 2: change of availability in reversible piston cylinder arrangement

Calculate the change of availability for the process described in Example 1. Compare the value obtained with the work terms evaluated in Example 1.

Solution

The appropriate definition of availability for System A is the non-flow availability function, defined in eqn (2.12a), and the maximum useful work is given by eqns (2.11) and (2.12b). Hence, w,,, is given by

w,,, = -Aa = - (a2 - a , ) = a , - a2 = 4 - u2 + P O ( V 1 - v2) - To(s, - s2)

Page 35: Advanced Thermodynamics for Engineers

Examples 19

The change of entropy, s1 - s2, can be evaluated from

dT dp d s = cP - - R -

T P which for a perfect gas becomes

Tl PI

T2 P2

s1 - s2 = cp In - - R In -

550 2

300 1 = 1.005 x In - - 0.287 x In - = 0.41024 kJ/kgK

Substituting gives

WuSe = C"(T1 - T2) + PO(V1 - v2) - TObI - s2)

1 x io5 io3 = (1.005 - 0.287)(550 - 300) + ~ (1 - 1.0909) x 0.7896

- 300 x 0.41024 = 179.5 - 7.17 - 123.07 = 49.25 kJ/kg

This answer is identical to the net useful work done by the system evaluated in part (vi) above; this is to be expected because the change of availability was described and defined as the maximum useful work that could be obtained from the system. Hence, the change of availability can be evaluated directly to give the maximum useful work output from a number of processes without having to evaluate the components of work output separately.

Example 3: availability of water vapour

Evaluate the specific availability of water vapour at 30 bar, 450°C if the surroundings are at p o = 1 bar; to = 35°C. Evaluate the maximum useful work that can be obtained from this vapour if it is expanded to (i) 20 bar, 250°C; (ii) the dead state.

Solution

From Rogers and Mayhew tables:

at p = 30 bar; t = 450°C u = 3020 kJ/kg; h = 3343 kJ/kg; v = 0.1078 m3/kg; s = 7.082 kJ/kg K

Hence the specific availability at these conditions is

a = u +pov - Tos = 3020 + 1 x lo5 x 0.1078/103 - 308 x 7.082 = 849.5 kJ/kg

(i) Change of availability if expanded to 20 bar, 250°C. From Rogers and Mayhew tables:

at p = 20 bar; t = 250°C u = 2681 kJ/kg; h = 2904 kJ/kg; v = 0.1 115 m3/kg; s = 6.547 kJ/kg K

Page 36: Advanced Thermodynamics for Engineers

20 Availability and exergy

The specific availability at these conditions is

a=u+p,v - T,s =2681 + I x lO5~O.1115/1O3-308x6.547 = 675.7 W/kg

Thus the maximum work that can be obtained by expanding the gas to these conditions is

w,, = a , - a2 = 849.5 - 675.7 = 173.8 kJ/kg

This could also be evaluated using eqn (2.12b), giving

Wmax = -‘a= h, + VI(PO-PI) - (h2 + ~ 2 ( ~ 0 - ~ 2 ) ) - TO(S, - ~ 2 )

= 3343 - 2904 + 0.1078 x (1 - 30) x lo2 - 0.11 15 x (1 - 20) x lo2 - 308 x (7.082 - 6.547)

=439-312.6+211.9- 164.78= 173.5 kJ/kg

The values obtained by the different approaches are the same to the accuracy of the figures in the tables. Change of availability if expanded to datum level at p o = 1 bar; to = 35°C (ii)

u = 146.6 kJ/kg; h = 146.6 kJ/kg; v = 0.1006 x

s = 0.5045 W/kg K m3/kg;

The specific availability at the dead state is

a = u +pov - T,s = 146.6 + 1 x lo5 x 0.001006/103 - 308 x 0.5045 = -8.7 M/kg

and the maximum work that can be obtained by expanding to the dead state is

w,, = a, - a2 = 849.5 - (-8.7) = 858.2 W/kg

Again, this could have been evaluated using eqn (2.12b) to give

Wm, = -Aa= A,+ V ~ ( P O -P I ) - (h2 + v ~ ( P o - P ~ ) ) - TO(s, - ~ 2 )

= 3343 - 146.6 + 0.1078 x (1 - 30) x lo2 - 0.001006 x (1 - 1) x lo2 - 308 x (7.082 - 0.5045)

= 3196.4 - 312.6 + 0 - 2025.9 = 857.9 kJ/kg

This shows that although the energy available in the system was 3020 kJ/kg (the internal energy) it is not possible to convert all this energy to work. It should be noted that this ‘energy’ is itself based on a datum of the triple point of water, but the dead state is above this value. However, even if the datum levels were reduced to the triple point the maximum useful work would still only be

w,, = -Aa = (3343 - 0) + 0.1078 x (0.00612 - 30) x lo2 - 273(7.082 - 0) = 1086.3 M/kg

It is also possible to evaluate the availability of a steady-flow system and this is defined in a similar manner to that used above, but in this case the reversible heat engine extracts

Page 37: Advanced Thermodynamics for Engineers

Irreversibility 21

energy from the flowing stream. If the kinetic and potential energies of the flowing stream are negligible compared with the thermal energy then the steady flow availability is

(2.13a) a , = h - Tos

The change in steady flow availability is

Aa, = af2 - a,, = h, - Tosz - ( h , - Tos,) (2.13b)

Thus, the maximum work that could be obtained from a steady-flow system is the change in flow availability, which for the previous example is

w,, = -Aa, = - ( a f 2 - u,,) = h, - Tos, - (h2 - Tos2) = (3343 - 146.6) - 308 x (7.082 - 0.5045) = 1170.4 kJ/kg

2.4 Available and non-available energy

If a certain portion of energy is available then obviously another part is unavailable - the unavailable part is that which must be thrown away. Consider Fig 2.4; this diagram indicates an internally reversible process from a to b. This can be considered to be made up of an infinite number of strips 1-m-n-4-1 where the temperature of energy transfer is essentially constant, i.e. TI = T4 = T. The energy transfer obeys

SQ SQo T To

-=-

where SQ = heat transferred to system and SQo = heat rejected from system, as in an engine (ER) undergoing an infinitesimal Camot cycle.

In reality SQ, is the minimum amount of heat that can be rejected because processes 1 to 2 and 3 to 4 are both isentropic, i.e. adiabatic and reversible.

Hence the amount of energy that must be rejected is

(2.14)

Note that the quantity of energy, SQ, can be written as a definite integral because the process is an isentropic (reversible) one. Then E,,, is the energy that is unavailable and is given by cdefc. The available energy on this diagram is given by abcda and is given by

E,, = Q - E,,, = Q - T d S (2.15)

where Q is defined by the area abfea.

2.5 Irreversibility

The concept of reversible engines has been introduced and these have operated on reversible cycles, e.g. isentropic and isothermal reversible processes. However, all real processes are irreversible and it is possible to obtain a measure of this irreversibility using the previous analysis. This will be illustrated by two examples: a turbine, which produces a work output; and a compressor, which absorbs a work input.

Page 38: Advanced Thermodynamics for Engineers

22 Availability and exergy

2" 5 3 t-"

b

TI = T4= T

C

f

Entropy, S

Fig. 2.4 Available and unavailable energy shown on a T-s diagram

Example 4: a turbine

An aircraft gas turbine with an isentropic efficiency of 85% receives hot gas from the combustion chamber at 10 bar and 1000°C. It expands this to the atmospheric pressure of 1 bar. If the temperature of the atmosphere is 20"C, determine (a) the change of availability of the working fluid, and the work done by the turbine ifthe expansion were isentropic. Then, for the actual turbine, determine (b) the change of availability and the work done, (c) the change of availability of the surroundings, and (d) the net loss of availability of the universe (Le. the irreversibility).

Assume that the specific heat at constant pressure, cp = 1.100 W/kg K, and that the ratio of specific heats, K = 1.35.

Solution

The processes involved are shown in Fig 2.5.

(a) Isentropic expansion From the steady flow energy equation the specific work done in the isentropic expansion is

wT li~en = cp(Tl - T 2 ' )

For the isentropic expansion, T2- = TI( p 2 / p I ) ( " - ' ) / " = 1273(1/10)0.35/'.35 = 700 - 9 8 K giving

wTlisen= 1.100 x (1273-700.8)=629.5 kJ/kg

The change in availability of the working fluid is given by eqn (2.13b), Aa = af2, - a,,, where

af2< = h,. - To+ a,, = h, - T,s,

For an isentropic change, s2- = s , , and hence Aa = h,. - h, = -wT lisen = -629.5 kJ/kg.

Page 39: Advanced Thermodynamics for Engineers

Irreversibility 23

The change of availability of the surroundings is 629.5 kJ/kg, and hence the change of availability of the universe is zero for this isentropic process. This means that the energy can be transferred between the system and the environment without any degradation, and the processes are reversible - this would be expected in the case of an isentropic process.

I

Entropy, S

Fig. 2.5 Turbine expansion on a T-s diagram

(b) Non-isentropic expansion If the isentropic efficiency of the turbine is 85%, i.e. vT = 0.85, then the turbine work is

wT = vTwT liscn = 0.85 x 629.5 = 535.0 kJ/kg

The temperature at the end of the expansion is

T2 = TI - wT/cP= 1273 - 535.0/1.100 = 786 K.

The change of availability is given by eqn (2.13b) as

Aa = af2 - a,,

= h2 - TOS~ - ( h , - TOsl)

= ~ ~ - ~ ~ - T O ( S ~ - S ~ ) = - W T - T O ( S ~ - S , )

The change of entropy for a perfect gas is given by

where, from the perfect gas law,

(IC- l ) ~ , - (1.35 - 1) x 1.100 R = - = 0.285 kJ/kg K

K 1.35

Page 40: Advanced Thermodynamics for Engineers

24 Availability and exergy

Thus the change of entropy during the expansion process is

786 1

1273 10 s- - s1 = 1.1 In - - 0.285 In - = -0.5304 + 0.6562 = 0.1258 kJ/kg K

Thus Aa = -535 - To(s2 - sl) = -535 - 293 x 0.1258 = -571.9 kJ/kg (c) The change of availability of the surroundings is equal to the work done, hence

(d) The irreversibility is the change of availability of the universe, which is the sum of the changes of availability of the system and its surroundings, i.e.

Aasurroundings = 535 k J / k

Aalmiverse = Aasystcm + Aasurroundings = -571.9 + 535.0 = -36.9 kJ/kg

This can also be calculated directly from an expression for irreversibility

I=To AS = 293 x 0.1258 = 36.86 kJ/kg

Note that the irreversibility is positive because it is defined as the loss of availability.

Example 5: an air compressor

A steady flow compressor for a gas turbine receives air at 1 bar and 15"C, which it compresses to 7 bar with an efficiency of 83%. Based on surroundings at 5 ° C determine (a) the change of availability and the work for isentropic compression. For the actual process evaluate (b) the change of availability and work done, (c) the change of availability of the surroundings and (d) the irreversibility.

Treat the gas as an ideal one, with the specific heat at constant pressure, c p = 1.004 kJ/ kg K, and the ratio of specific heats, 1c = 1.4.

Solution

The processes are shown in Fig 2.6.

/ p 2 = 7 b a r

bar

~~

Entropy, S

Fig. 2.6 Compression process on a 7'-s diagram

Page 41: Advanced Thermodynamics for Engineers

Graphical representation of available energy 25

(a) Isentropic compression: (K - l ) / K

T,. = T ( :) = 288 x 7°'286 = 502.4 K

The work done can be calculated from the steady flow energy equation, giving

wC lisen = -Ah = -215.3 kJ/kg

The change of availability is given by eqn (2.13b) as

Aa, = auP - a,, = h2, - h, - T,(s,. - s l ) = Ah - T A S

If the process is isentropic, AS = 0, and then

Aaf = cp(T2. - T , ) = 1.004 x (502.4 - 288) = 215.3 kJ/kg

(b) The actual work done, wc = wc Iisen/vC = -215.3/0.83 = -259.4 kJ/kg. Hence h, = h, - wc = 1.004 x 288 - (-259.4) = 548.6 kJ/kg, and the temperature at 2, T2, is given by T2 = h2/c , = 546.4 K.

T2 P2

Ti P1 Asl2 = c p In - - R In - = 0.641 1 - 0.5565 = 0.08639 kJ/kg K

Hence Aa, = af2 - a,, = A h - T& = 259.4 - 278 x 0.08639 = 235.4 kJ/kg

There has been a greater increase in availability than in the case of reversible compression. This reflects the higher temperature achieved during irreversible compression. The change of availability of the surroundings is equal to the enthalpy change of the system, i.e. AaSw = -259.4 kJ/kg. The process an adiabatic one, i.e. there is no heat loss or addition. Hence the irreversibility of the process is given by

(c)

(d)

I = To AS = 278 x 0.08639 = 24.02 kJ/kg

This is the negative of the change of availability of the universe (system + surroundings) which is given by

A a ~ " = Aasystcm + Aasmundlngs = 235.4 - 259.4 = -24.0 kJ/kg

Hence, while the available energy in the fluid has been increased by the work done on it, the change is less than the work done. This means that, even if the energy in the gas after compression is passed through a reversible heat engine, it will not be possible to produce as much work as was required to compress the gas. Hence, the quality of the energy in the universe has been reduced, even though the quantity of energy has remained constant.

2.6 Graphical representation of available energy, and irreversibility

Consider the energy transfer from a high temperature reservoir at TH through a heat engine (not necessarily reversible), as shown in Fig 2.7.

Page 42: Advanced Thermodynamics for Engineers

26 Availability and exergy

c

Fig. 2.7 Representation of available energy, and irreversibility

The available energy flow from the hot reservoir is

EH=QH-ToASH

The work done by the engine is

W = QH - Qo

The total change of entropy of the universe is

Hence the energy which is unavailable due to irreversibility is defined by

Ei,,ev= E,- W=QH- To ASH- W

= QH - To AsH(QH - Qo) = Qo - To ASH = To(AS0 - ASH)

In the case of a reversible engine c AS = 0 because entropy flow is conserved, i.e.

QH Qo - = -

TH TO

(2.16)

(2.17)

(2.18)

(2.19)

Hence the unavailable energy for a reversible engine is To ASH while the irreversibility is zero. However, for all other engines it is non-zero. The available energy is depicted in Fig 2.7 by the area marked ‘A’, while the energy ‘lost’ due to irreversibility is denoted ‘I’ and is defined

Eimv = To(AS0 - ASH) (2.20)

Page 43: Advanced Thermodynamics for Engineers

Availability balance for a closed system 27

2.7 Availability balance for a closed system

The approaches derived previously work very well when it is possible to define the changes occurring inside the system. However, it is not always possible to do this and it is useful to derive a method for evaluating the change of availability from ‘external’ parameters. This can be done in the following way for a closed system.

If a closed system goes from state 1 to state 2 by executing a process then the changes in that system are:

2 2 from the First Law: U2 - U , = ( S Q - SW) = I SQ - W (2.21)

(2.22)

1

2 SQ from the Second Law: S2 - S , = I - + u l T

where u is the internal irreversibility of the system, and T is the temperature at which the heat transfer interactions with the system occur (see Fig 2.2(a)). Equations (2.21) and (2.22) can be written in terms of availability (see eqn (2.10)), for a system which can change its volume during the process, as

2 2 S Q = 1, SQ - W + po( V2 - Vi ) - To 1, - - Too

T (2.23)

Equation (2.23) can be rearranged to give

(2.24)

The change in specific availability is given by

a2 - a , = u2 - u , - To(sz - sl) +po(v2 - v,)

= 1,’ (1 - :)dq - w + p0(v2 - v l ) - T,a, (2.25)

where q, w and u, are the values of Q, Wand u per unit mass.

examples. The significance of eqn (2.24) can be examined by means of a couple of simple

Example 6: change in availability for a closed system

A steel casting weighing 20 kg is removed from a furnace at a temperature of 800°C and heat treated by quenching in a bath containing 500 kg water at 20°C. Calculate the change in availability of the universe due to this operation. The specific heat of the water is 4.18 kJ/kg K, and that of steel is 0.42 kJ/kg K. Assume that the bath of water is rigid and perfectly insulated from the surroundings after the casting has been dropped in, and take the datum temperature and pressure as 20°C and 1 bar respectively.

Page 44: Advanced Thermodynamics for Engineers

28 Availability and exergy

Solution

The process can be considered to be a closed system if it is analysed after the casting has been introduced to the bath of water. Hence eqn (2.24) can be applied:

A, - A, = u, - u, - &(S, - s,) +Po(& - y ) = J 1 - - SQ - w + - v,) - 1 z ( :.)

In this case, if the combined system is considered, SQ = 0, W = 0 and po(Vz - VI) = 0 because the system is adiabatic and constant volume. Thus

A , - A , = -TOO

The irreversibility can be calculated in the following manner:

1. final temperature of system

m,c,T, + mwcwTw

m,c, + m,c,

change of entropy of casting

20 x 0.42 x 1073 + 500 x 4.18 x 293

0.42 x 20 + 500 x 4.18 - T, = -

= 296.12 K

2.

= -10.815 kJ/K

3. change of entropy of water

Tf 296.12 = mwcw In - = 500 x 4.18 x In ~

TW 293 = 22.154 kJ/K

4. change of entropy of system (and universe)

SS = 6S, + SS, = -10.815 + 22.154 = 11.339 U/K 5. change of availability

A , - A , = - T o a = - 2 9 3 ~ 11.34~-3323k.l

This solution indicates that the universe is less able to do work after the quenching of the casting than it was before the casting was quenched. The loss of availability was analysed above by considering the irreversibility associated with the transfer of energy between the casting and the water. It could have been analysed in a different manner by taking explicit note of the work which could be achieved from each part of the composite system before and after the heat transfer had taken place. This will now be done.

Work available by transferring energy from the casting to the environment through a reversible heat engine before the process

1.

W , = 1 (1 - :) dQ = 1; (1 - :)mccc dT = m,c,[T - To In TITc TO

T ,

(293 - 1073) - 293 x In

Page 45: Advanced Thermodynamics for Engineers

Availability balance for a closed system 29

2. Work available by transfening energy from the water bath to the environment through a reversible heat engine after the process

To W , = (1 - $!) dQ = 1; (1 - :)mwcw dT = rnwcw[T - To In TITf

(293 - 296.12) - 293 x In

3. Change of availability of universe is

A , - A , = -W= ~(3357.3-34.47)= -3323 kJ

This is the same solution as obtained by considering the irreversibility. Equation (2.24) can be considered to be made up of a number of terms, as shown below:

Too - availability transfer availability destruction accompanying work due to irrevenibilities availability transfer

accompanying heat transfer

I = II2 (1 - :)dQ - *W + po(V , - V1! - Y availabi]ity transfer availability destruction accompanying work due to irreversibilities

availability transfer

(2.26)

accompanying heat transfer

It is also possible to write eqn (2.26) in the form of a rate equation, in which case the rate of change of availability is

dV

dt - w + p , - - T0a - . . , avaiiabilitydestruction

avalability transfer due to irreversibilities

dt - availability transfer

accompanying heat transfer accompanying work

dV I = (l-?)Q - w + p o - dt - rusliirhiiim v ~esmction

".-.l" ...., I P availability transfer due to I availability transfer

accompanying heat transfer accompanying work

nibilities

Equations (2.26) and (2.27) can be written in terms of specific availability to give

a2 - a , = (1 - :)dq - w +po(v, - v l > - Toom

= j12 (1 - ?)dq - w + po(v, - V I > - i

(2.27)

(2.28)

dv . dt (2.29)

da dv

dt dt

Page 46: Advanced Thermodynamics for Engineers

30 Availability und exergy

Example 7: change of availability for a closed system in which the volume changes

An internal combustion engine operates on the Otto cycle (with combustion at constant volume) and has the parameters defined in Table 2.1. The data in this example is based on Heywood (1988).

Table 2.1 Operating parameters for Otto cycle

Compression ratio, rc Calorific value of fuel, Qb/(kJ/kg) Ratio, x, = AAR/Q: Pressure at start of compression, p,/(bar) Temperature at start of compression, T,/(OC) Air/fuel ratio, E

Specific heat of air at constant volume, c,/(kJ/kg K ) Ratio of specific heats, K Temperature of surroundings, To/ (K) Pressure of surroundings, po/(bar)

12 44 000 1.0286 1 .o 60 15.39 0.946 1.30 300 1 .o

Calculate the variation in availability of the gases in the cylinder throughout the cycle from the start of compression to the end of expansion. Assume the compression and expansion processes are adiabatic.

Solution This example introduces two new concepts: 0

0

the effect of change of volume on the availability of the system; the effect of ‘combustion’ on the availability of the system.

Equation (2.24) contains a term which takes into account the change of volume ( p o ( V z - V , ) ) , but it does not contain a term for the change in availability which occurs due to ‘combustion’.

Equation (2.24) gives the change of availability of a system of constant composition as an extensive property. This can be modified to allow for combustion by the addition of a term for the availability of reaction. This is defined as

where T, is the temperature at which the energies of reaction are evaluated. Hence

(2.31)

Equation (2.32) can be generalised to give the value at state i relative to that at state 0, resulting in

Ai-Ao=AAR+ (Uj- U0)- To(Sj-So)+po(Vj-Vo) = m f A A R + m [ ( u j - u o ) - T o ( ~ i - ~ o ) + p o ( ~ i - ~ o ) ]

where m, = mass of fuel (2.33)

and m = total mass of mixture = m, + mair

Page 47: Advanced Thermodynamics for Engineers

Availability balance for a closed system 3 1

The specific availability, based on the total mass of mixture, is

(2.34)

where &=overall air-fuel ratio. Examples of the availability of reaction are given in section 2.9.2. The availability of reaction can be related to the internal energy of reaction by

which gives

(2.35)

(2.36)

If the energy of reaction per unit mass of mixture is written q* = ( Q , ) , / ( E + 1) then

(2.37)

Considering the individual terms in eqn (2.37)

Po

Ti * a , - a , c,T,

Hence, - - 1 +-- K h - - ( K - l ) h - 4* --[(z-) 4* zo [ To

i- ( K - I)(: - I)]

(2.38)

(2.39)

(2.40)

(2.41)

Equation (2.41) can be applied around the cycle to evaluate the availability relative to that at the datum state. The values at the state points of the Otto cycle are shown in Tables 2.2 and 2.3, and the variation of availability as a function of crankangle and volume during the cycle (calculated by applying eqn (2.41) in a step-by-step manner) is shown in Figs 2.8 and 2.9 respectively.

Page 48: Advanced Thermodynamics for Engineers

32 Availability and exergy

Table 2.2 State points around Otto cycle

0 1 1 300 1.0286 1 1 . 1 1 1 333 1.029375 0.000775 2 0.0925 25.28923 701.7762 1.127041 0.097666 3 0.0925 127.5691 3540.043 0.955991 -0.17105 4 1 . 1 1 5.044404 1679.787 0.332836 -0.62316

Table 2.3 Terms in eqn (2.41) evaluated around cycle

0 0 9.73 1469 0 0 1.0286 1 0.11 9.731469 0.033 0.135668 1.029375 0.000775 2 1.339254 9.731469 -0.27225 0.135668 1.127041 0.097666 3 10.80014 9.731469 -0.27225 1.753948 0.955991 -0.17105 4 4.599289 9.731469 0.033 1.753948 0.332836 -0.62316

The value of availability of the charge at state 0 is based on the ratio of the availability of combustion of octane to the heat of reaction of octane, and this is calculated in section 2.9.2. The calculation of most of the other points on the cycle is straightforward, but it is worthwhile considering what happens during the combustion process. In the Otto cycle combustion takes place instantaneously at top dead centre (tdc), and the volume remains constant. This means that when eqn (2.41) is used to consider the effect of adiabatic, constant volume combustion occurring between points 2 and 3, this gives

a3-ao a2-a .

4* 4*

- I)-{(: - 1). $} - {Kln - T3 - (K- 1)ln T2

change of availability due to combustion as fuel changes from reactants to products, = 0

change of availability of gases due to change of enmpy of gases

Hence, the availability at point 3 is

-- +- - - K l n - - ( K - l ) h - - a 3 - ~ 0 a 2 - a ,

4* 4* cvTo 4* [ i :: P2 p311

-- (2.43)

Equation (2.43) is the change in entropy of the working fluid which is brought about by combustion, and since the entropy of the gases increases due to combustion this term reduces the availability of the gas.

Figures 2.8 and 2.9 show how the non-dimensional availability of the charge varies around the cycle. This example is based on a cycle with a premixed charge, i.e. the fuel and

Page 49: Advanced Thermodynamics for Engineers

Availability balance for a closed system 33

1.20

*m 1.10 . 1.00 m

x .z 0.90

2 0.80

2 0.70

- .e

4 .+

m 0.60 - -_

d 0.50

.+ 0 cj 0.40

.$ 0.30

b 0.20

g 0.10

0.00

A v a i l a b i l i t y r e c e i v e d d u r i n g c o m p r e s s i o n

A v a i l a b l i t y d e s t r o y e d d u r i n g c o m b u s t i o n *

A v a i l a b i l i t y d e l i v e r e d a s work d u r i n g

e x p a n s i o n

A v a i l a b i l i t y in c h a r g e

.180 -135 -90 -45 0 45 90 135 180

Crankangle / (deg)

Fig. 2.8 Variation of the non-dimensional availability with crank angle for an Otto cycle

* 0- m . x c) .+ - .+ n m

m .+ > m

1.10

1.00

0.90

0.80

0.70

0.60

0.50

0.40

0.30

0.20

0.10

0.00

-~ A v a i l a b i l i t y r e c e i v e d d u r i n g c o m p r e s s i o n

A v a i l a b i l i t y d e s t r o y e d d u r i n g c o m b u s t i o n

A v a i l a b i l i t y d e l i v e r e d a s work d u r i n g e x p a n s i o n IY

0 1 2 3 4 5 6 7 8 9 1 0 1 1 1 2

VI vc Fig. 2.9 Variation of non-dimensional availability with volume for an Otto cycle

air are induced into the engine through the inlet valve, and the availability of the charge is greater at the beginning of the cycle than the end because the fuel contains availability whch is released during the chemical reaction. This is similar to the energy contained in the reactants in a simple combustion process and released as the internal energy (or enthalpy) of reaction of the fuel. Considering Fig 2.8, it can be seen that the availability of the charge

Page 50: Advanced Thermodynamics for Engineers

34 Availability and exergy

increases during the compression stroke (- 180" crankangle to 0" crankangle) because of the work done by the piston (which is isentropic in this ideal cycle case). The 'combustion' process takes place instantaneously at 0" crankangle and the availability contained in the fuel in the form of the chemical bonds (see Chapter 11) is released and converted, at constant volume, to thermal energy. The effect of this is to introduce an irreversibility defined by the change of entropy of the gas (eqn (2.42)), and this causes a loss of availability. The expansion process (which is again isentropic for this ideal cycle) causes a reduction of availability in the charge as it is converted into expansion work.

The availability of the working fluid at the end of the cycle, 4, is a measure of the work that could be obtained from the charge after the cycle in the cylinder has been completed. This availability can be obtained if the working fluid is taken down to the dead state by reversible processes. Turbochargers are used to convert part of this availability into work but cannot convert it all because they can only take the state of the fluid down to the pressure of the dead state, po. If the full availability is to be converted it is necessary to use a turbine (e.g. a turbocharger) and then a 'bottoming' cycle (e.g. a Rankine cycle based on a suitable working fluid) to take the exhaust gas down to the dead state temperature, To.

The non-dimensional availability, a/q* , is a measure of the work that can be obtained from the cycle. Hence, the total 'non-dimensional work' that can be achieved is

da, da, da,, * + * = 1.029375 - 0.332836 - 0.17105 = 0.5255

4* 4 4 -- net work from availability

availability generated

This is the ratio of the net work output to the energy supplied (q * ) , and is equal to the thermal efficiency of the engine cycle. The thermal efficiency of this Otto cycle, with a compression ratio of 12, is

1 ?jJoao = 1 - - - - 0.5255 p- 1)

Thus, the availability approach has given the same result as the basic equation. Where the availability approach shows its strength is in analysing the effects of finite combustion rates and heat transfer in real cycles; this has been discussed by Patterson and Van Wylen (1964).

2.8 Availability balance for an open system The availability balance of a closed system was derived in section 2.7. A similar approach can be used to evaluate the change of availability in an open system. The unsteady flow energy equation is

c Q - W = (:) + h e ( h e + V:/2 + gz,) - h i ( h i + V;/2 + gzi) (2.44) i cv e i

where C j Q is the sum of heat transfers to the control volume over the total number of heat transfer processes, j , and the summations of the inflows and outflows take into account all the inflows and outflows. If the changes in kinetic and potential energies are small then this equation can be reduced to

cQ - W = (:) + x h e h e - 1 h i h i i CY e i

(2.45)

Page 51: Advanced Thermodynamics for Engineers

Availability balance for an open system 35

It is also possible to relate the rate of change of entropy to the heat transfers from the system, and the irreversibilities in the system as

s=-- dt j T e I

(2.46)

If the kinetic and potential energies of the control volume do not change then the internal energy in the control volume, (E)w, can be replaced by the intrinsic internal energy of the control volume, (U)cv. The availability of the control volume is defined by eqn (2.10), and hence

Substituting eqns (2.46) and (2.47) into eqn (2.45) gives

= F (1 - :)Q - (W -po z) + Cki(hi - Tosi) dt i

- @(he - T ~ s , ) - To~j., e

The terms h - Tos are the flow availability, a,, and hence eqn (2.48) becomes

(2.47)

(2.48)

(2.49)

where Zc, is the irreversibility in the control volume. Equation (2.49) is the unsteady flow availability equation, which is the availability

equivalent of the unsteady flow energy equation. Many of the processes considered in engineering are steady state ones, which means that the conditions in the control volume do not change with time. This means that dA,,/dt = 0 and dV/dt = 0, and then eqn (2.49) can be simplified to

o = C 1 - - Q - W + Chiaf, - Ckeafe - T~$, , i (:) i e

&,af, - C keafe - i,, e

Example 8: steady flow availability

(2.50)

Superheated steam at 30 bar and 250°C flows through a throttle with a downstream pressure of 5 bar. Calculate the change in flow availability across the throttle, neglecting the kinetic and potential terms, if the dead state condition is To = 25°C and po = 1 bar.

Page 52: Advanced Thermodynamics for Engineers

36 Availability and exergy

Solution

A throttle does not produce any work, and it can be assumed that the process is adiabatic, i.e. Q = 0, W = 0. Hence, eqn (2.50) becomes, taking into account the fact that there is only one inlet and outlet,

0 = mlaf, - meafe - I , ,

The conditions at inlet, i , are h, = 2858 kJ/kg, s, = 6.289 kJ/kg K; and the conditions at exhaust, e, are he = 2858 kJ /kg, s, = 7.0650 W/kg K.

Hence the irreversibility per unit mass flow is

= (2858 - 298 x 6.289) - (2858 - 298 x 7.0650) = 231.28 kJ/kg

The significance of this result is that although energy is conserved in the flow through the throttle the ability of the fluid to do work is reduced by the irreversibility. In this case, because the enthalpy does not change across the throttle, the irreversibility could have been evaluated by

T0(s2 - s,) = 298 x (7.065 - 6.289) = 231.28 kJ/kg

2.9 Exergy

Exergy is basically the available energy based on datum conditions at a dead state; it was introduced in Example 3. An obvious datum to be used in most calculations is the ambient condition, say po,T,. This datum condition can be referred to as the dead state and the system reaches this when it is in thermal and mechanical equilibrium with it. A state of thermal and mechanical equilibrium is reached when both the temperature and pressure of the system are equal to those of the dead state. The term exergy was proposed by Rant in 1956, and similar functions had previously been defined by Gibbs and Keenan.

Exergy will be given the symbol B (sometimes it is given the symbol S ) and specific exergy will be denoted b (or 6). The exergy of a system at state 1 is defined by

B, = A , - A0 (2.51)

where A , is the available energy at state 1 , and A , is the available energy at the dead state.

If a system undergoes a process between states 1 and 2 the maximum useful work, or available energy, that may be obtained from it is given by

W = A , - A , = ( A , - A, ) - (A2 - A,) = Bl - B, (2.52)

Hence the maximum work, or heat transfer, that can be obtained as a system changes between two states is defined as the difference in exergy of those two states.

Exergy is very similar to available energy and is a quasi- or pseudo-property. This is because it is not defined solely by the state of the system but also by the datum, or dead, state that is used.

A number of examples of the use of exergy will be given.

Page 53: Advanced Thermodynamics for Engineers

Exergy 37

2.9.1 HEAT TRANSFER

It is possible for energy to be transferred from one body to another without any loss of available energy. This occurs when the heat or energy transfer is ideal or reversible. No real heat transfer process will exhibit such perfection and exergy can be used to show the best way of optimising the effectiveness of such an energy transfer. All actual heat transfer processes are irreversible and the irreversibility results in a loss of exergy.

It should be noted in this section that even though the heat transfer processes for each of the systems are internally reversible, they might also be externally irreversible.

Ideal, reversible heat transfer

Ideal reversible heat transfer can be approached in a counterflow heat exchanger. In this type of device the temperature difference between the two streams is kept to a minimum, because the hot ‘source’ fluid on entering the heat exchanger is in closest contact to the ‘sink’ fluid which is leaving the device, and vice versa. The processes involved are depicted by two almost coincident lines from 1 to 2 in Fig 2.10.

In this ideal process it will be assumed that, at all times, the fluid receiving the heat is at temperature T , while the temperature of the source of heat is at all times at temperature T + 6 T , i.e. TI, is 6T less than T,, etc. From the First Law of Thermodynamics it is obvious that, if the boundaries of the control volume are insulated from the surroundings, the energy transferred from the hot stream must be equal to the energy received by the cold stream. This means that the areas under the curves in Fig 2.10 must be equal: in this case they are identical. The exergy change of the hot stream is then given by

ABh=B2h-Blh= ( A 2 h - A O ) - ( A l h - A O )

=A, , -AIh= ( U + p o V - T0S),,- ( U + P ~ V - T , S ) , ~

= ( U 2 h - U l h > - - S l h ) + P 0 ( V 2 h - v l h ) (2.53)

If it is assumed that V I , = V2, then

= ( U 2 h - Ulh) - - ’1,) (2.54)

In eqn (2.54), U,, - Ulh= heat transferred from the hot stream, and S,, - SI,= the change of entropy of the hot stream. In Fig 2.10, UZh- U,,= area lh-2h-6-5-lh and To(S, - SI,,) =area 3-4-6-5-3. In the case of the hot stream the change of energy and the change of entropy will both be negative, and because (UZh- Ulh)> To(&,- Sib) the change of exergy will also be negative.

A similar analysis can be done for the cold stream, and this gives

ABc = B , - BI, = (AZc - A,) - ( A l c - A , )

= (U2c - Ulc) - TO(S2, - SI,) (2.55)

In this case U,, - U,, = heat transferred to the cold stream (area lc-2c-5-6-lc), and this will be positive. The term To(S,, - S,,) = unavailable energy of the cold stream (area 4-3- 5-6-4), and this will also be positive because of the increase of entropy of the cold stream. Figure 2.10 shows that the heating and cooling processes are effectively identical on the T-s diagram and only differ in their direction. Hence, the changes of energy of the hot and cold streams are equal (and opposite), and also the the unavailable energies are equal (and opposite) for both streams. This means that for an ideal counterflow heat exchanger, in which the heat transfer takes place across an infinitesimal temperature difference (i.e.

Page 54: Advanced Thermodynamics for Engineers

38 Availability and exergy

reversible heat transfer), the loss of exergy of the hot stream is equal to the gain of exergy of the cold stream. The change of exergy of the universe is zero, and the process is reversible.

lC/

lh 2c

3

5

Entropy, S

Fig. 2.10 Reversible heat transfer in a counterflow heat exchanger: (a) schematic diagram of heat exchanger; (b) processes shown on a T-s diagram

If the heat exchanger in Fig 2.10 were not reversible, i.e. there is a significant temperature difference between the hot and cold stream, then the T-s diagram would be like that shown in Fig 2.1 1. Figure 2.11 (a) shows the processes for both the hot and cold streams on the same T-s diagram, while Figs 2.11 (b) and 2.1 1 (c) show individual T-s diagrams for the hot and cold stream respectively. The first point to recognise is that the energies transferred between two streams (areas lh-2h-5h-6h-lh and lc-2c-5c-6c-lc) are equal (and opposite). The change of exergy for the hot stream is denoted by the area lh- 2h-3h-4h- 1 h, while the unavailable energy is given by 3h-5h-6h-4h-3h. Similar quantities for the cold streams are lc-2c-3c-4c-lc and 3c-5c-6c-4c-3c. It is obvious from Fig 2.11 (a) that the unavailable energy of the cold stream is greater than that for the hot stream because To(S, - SI,) > To(Slh - &h). This means that the exergy gained by the cold stream is less than that lost by the hot stream. For the hot stream

ABh = (UZh - Ulh) - - Slh) (2.56)

while for the cold stream

ABC = (Uzc - Ulc) - TO(S2 - SIC) (2.57)

The change of exergy of the universe is given by

= ABh + ABc = (Uzh - UIh) - To(S2, - Sib) (U, - Uic) - SIC) = - + S2c - $1,) (2.58)

The value of AB^" c 0 because the entropy change of the cold stream is greater than that of the hot stream. Hence, in this irreversible heat transfer device there has been a loss of exergy in the universe, and energy has been degraded. This means that the maximum useful work available from the cold stream is less than that available from the hot stream, even though the energy contents of the two streams are the same.

Page 55: Advanced Thermodynamics for Engineers

Exergy 39

2h

I T

l h

To

T

Entropy, S (4 To

/ Hot stream

2h

3h

Sh

@)

h

4h

6h S -

2c

3c

5c - S

Fig. 2.11 T-s diagrams for a counterflow heat exchanger with finite temperature difference between the streams

This result has a further significance, which is that, to obtain the maximum transference of exergy, it is important to reduce the value of To ( S , - SJ. If the total amount of energy being transferred is kept constant then the loss of exergy can be minimis4 by increasing the temperature at which the heat is transferred. This is effectively shown in Fig 2.11, where the high temperature stream can be equivalent to a high temperature source of heat, while the low temperature stream is equivalent to a lower temperature source of heat. The quality of the heat (energy) in the high temperature source is better than that in the low temperature one.

Irreversible heat transfer

The processes depicted in Fig 2.12 are those of an infinite heat source transferring energy to a finite sink. The temperature of the source remains constant at TI but that of the sink changes from T, to Td. The energy received by the sink will be equal to that lost by the source if the two systems are isolated from the surroundings. The process undergone by the source is one of decreasing entropy while that for the sink is one of increasing entropy. Hence, in Fig 2.12, the areas 1-2-8-9-1 and 5-6-10-8-5 are equal. This means that areas

(a+ b + c ) = (b+ c + d + e )

By definition the exergy change of the source is given by

(2.59)

Page 56: Advanced Thermodynamics for Engineers

40 Availability and exergy

Entropy, S

Fig. 2.12 Irreversible heat transfer from an infinite reservoir to a finite sink

This equation assumes that the heat transfer takes place at constant volume. In a similar

(2.60)

Since the source and sink are isolated from the surroundings (the remainder of the universe), then the entropy change of the universe is

way the exergy change of the sink is given by

ABs,* = B, - B, = ( U , - U,) - TO(S6 - S,)

AB-v = AB,, + AB,,,

= ( = T d S , - S,) (2.61)

The term Tn(Si - S,) is depicted by the area marked ‘e’ in Fig 2.12. Since S6 is greater than SI then the exergy of the universe (that is, its ability to do work) has decreased by this amount. Thus, whilst the energy of the universe has remained constant, the quality of that energy has declined. This is true of all processes which take place irreversibly; that is, all real processes.

- ui ) - Tn(S, - Si ) + [ ( u, - u5) - To(S, - S5) I

2.9.2

Combustion processes are a good example of irreversible change. In a combustion process the fuel, usually a hydrocarbon, is oxidised using an oxidant, usually air. The structure of the hydrocarbon is broken down as the bonds between the carbon and hydrogen atoms are broken and new bonds are formed to create carbon dioxide, carbon monoxide and water vapour. These processes are basically irreversible because they cannot be made to go in the opposite direction by the addition of a very small amount of energy. This seems to suggest that exergy of the universe is decreased by the combustion of hydrocarbon fuels. The following section describes how combustion can be considered using an exergy approach.

Consider a constant pressure combustion process. When the system is in equilibrium with its surroundings the exergy of component i is:

(2.62)

EXERGY APPLIED TO COMBUSTION PROCESS

b, = ( A , - A,) - m s , - $01

Page 57: Advanced Thermodynamics for Engineers

Exergy 41

Exergy of reaction of water

Applying eqn (2.62) to the simple reaction

2H,(g) + O,(g>+2H,O(g)

AB = B2- B l = Bp- BR=C(br)p-C(bl)R

where suffix R indicates reactants and suffix P indicates products. Thus

AB = (hy)Hzo - To(SH~O - SH* - 0.5S02)

(2.63)

(2.64)

(2.65)

Substituting values, for To = 25"C, which is also the standard temperature for evaluating the enthalpy of reaction, gives

AB = -241820 - 298 x (188.71 - 130.57 - 0.5 x 205.04) = -228594.76 kJ/kmol K (2.66)

This means that the ability of the fuel to do work is 5.5% less than the original enthalpy of formation of the 'fuel', and hence 94.5% of the energy defined by the enthalpy of formation is the maximum energy that can be obtained from it.

Exergy of reaction of methane (CH,)

The equation for combustion of methane is

CH, + 20,+ CO, + 2H20 Hence

AB = C(bj)p- C(bj)R = (AHRICH, - T O ( S C O ~ + 2SH20 - SC& - 2S0,) = -804.6 x lo3 - 298 x (214.07 + 2 x 188.16 - 182.73 - 2 x 204.65) = -804.1 x 103 u/km0i

(2.67)

In this case the exergy of reaction is almost equal to the enthalpy of reaction; this occurs because the entropies of the reactants and products are almost equal.

Exergy of reaction of octane (C,H,,)

The equation for combustion of octane is

C,H,, + 12.5O,-+8CO2 + 9H20 (2.68)

Assume that the enthalpy of reaction of octane is -5074.6 x lo3 kJ/kmol, and that the entropy of octane at 298 K is 360 kJ/kmol K. Then the exergy of reaction is

AB = C(WP - C ( b , ) , = (AHR)c,H,, - To(8~c0, + 9sHzo - $C8Hl8- 12.589) = -5074.6 x lo3 - 298 x (8 x 214.07 + 9 x 188.16 - 360 - 12.5 x 204.65) = -5219.9 x lo3 kJ/kmol

In this case the exergy of reaction is greater than the enthalpy of reaction by 2.86%.

Page 58: Advanced Thermodynamics for Engineers

42 Availability and exergy

2.10 The variation of flow exergy for a perfect gas

This derivation is based on Haywood (1980). The definition of exergy is, from eqn (2.51),

B l = A , -A0 (2.51)

while that for the exergy of a flowing gas is, by comparison with eqn (2.13b) for the availability of a flowing gas,

(2.69) Bf, = A f , - A,, = w 1 - TOS,) - (Ho - TOSO) Equation (2.69) can be expanded to give the specific flow exergy as

b f , = a,, - afo = (hl - To$,) - (ho - Toso) (2.70)

Now, for a perfect gas,

h = cpT

and the change of entropy

T P

TO Po s1 - so = cp In - - R In -

Hence, the flow exergy for a perfect gas can be written

(2.71)

(2.72)

Po Po

T b n = ( h , - h o ) - T

(2.73)

The flow exergy can be non-dimensionalised by dividing by the enthalpy at the dead state temperature, To, to give

(2.74)

Equation (2.74) has been evaluated for a range of temperature ratios and pressure ratios, and the variation of exergy is shown in Fig 2.13.

m m e, c 0 e

.e

3 n E .-

6

Non-dimensional temperature (TmJ

Fig. 2.13 Variation of exergy with temperature and pressure for a perfect gas

Page 59: Advanced Thermodynamics for Engineers

Problems 43

Also shown in Fig 2.13, as a straight line, is the variation of enthalpy for the parameters shown. It can be seen that the exergy is sometimes bigger than the enthalpy, and vice versa. This is because the enthalpy is purely a measure of the thermodynamic energy of the gas relative to the datum temperature. When the dimensionless temperature T/To< 1.0 the enthalpy is negative, and when T/To > 1.0 the enthalpy is positive. This simply means that the thermodynamic energy can be greater than or less than the datum value. While the enthalpy varies montonically with the non-dimensional temperature, the exergy does not. The reason for this is that the exergy term, say at p / p o = 1, is given by

For

(2.75)

- > l , T ( k - l ) - h - > l T

TO TO

and also for

The physical significance of this is that the system can produce work output as long as there is a temperature difference between the system and the dead state: the sign of the temperature difference does not matter.

2.1 1 Concluding remarks

This chapter has introduced the concept of the quality of energy through the quasi- properties availability and exergy. It has been shown that energy available at a high temperature has better quality than that at low temperature. The effect of irreversibilities on the quality of energy has been considered, and while the energy of the universe might be considered to remain constant, the quality of that energy will tend to decrease.

It was also shown that the irreversibility of processes can be calculated, and it is this area that should be tackled by engineers to improve the efficiency of energy utilisation in the world.

PROBLEMS 1 A piston-cylinder assembly contains 3 kg of air at 15 bar and 620 K. The environment

is at a pressure of 1 bar and 300 K. The air is expanded in a fully reversible adiabatic process to a pressure of 5.5 bar. Calculate the useful work which can be obtained from this process. Also calculate the maximum useful work which can be obtained from the gas in (a) the initial state, and (b) the final state.

Assume that the specific heats for the gas are cp = 1.005 H/kg K, and C, = 0.718 kJ/kg K.

[295.5 kJ; 509.6 kJ; 214.1 kJ]

2 Air passes slowly through a rigid control volume C, as shown in Fig P2.2(a), in a hypothetical, fully reversible, steady-flow process between specified stable end states

Page 60: Advanced Thermodynamics for Engineers

44 Availability and exergy

1 and 2 in the presence of an environment at temperature To and pressure p o . States 1 and 2, and also the values of To and p o , are defined below:

T ,=550K and p 1 = 2 bar Tz=To=300K and p 2 = p o = 1 bar

The air may be treated as a perfect gas, with cp = 1.005 kJ/kg K and c, = 0.718 kJ/

The air follows path 1-a-2 see Fig P2.2(b); calculate the following work output quantities in each of the sub-processes 1-a and a-2:

The direct shaft work, W,, coming from the control volume as the air passes through the given sub-process. The external work, We, produced by any auxiliary cyclic devices required to ensure external reversibility in that exchange with the environment during the given sub-process.

kg K. (a)

(i)

(ii)

PI 7

Svstem B / - C - - - - - - - - - - ---

r System A '. c

/

I I '\ Exitstate,2 I Rigid control

P, I// p,

I Entropy, S

Fig. P2.2

Page 61: Advanced Thermodynamics for Engineers

Problems 45

(iii) The gross work Wg. (iv) The total useful shaft work Wxc.

[66.03; 62.2; 128.23; 128.23 (kJ/kg)]

Calculate the above work output quantities for each of the sub-processes 1-b and b-2 when the air follows the alternative path 1-b-2.

[109.41; 18.77; 128.18; 128.18 (kJ/kg)]

(b)

3 Confirm that item (iv) in Q.2. is equal to (b , - b2) , where b = h - Tos, the specific steady-flow availability function. (Note that, since T, = To and p 2 = p o , state 2 is the dead state, so that item (iv) is in this case equal to the steady-flow exergy of unit mass of air in state 1 for an environment at To = 300 K and p o = 1 bar.)

[128.2 kJ/kg]

4 A mass of 0.008 kg of helium is contained in a piston-cylinder unit at 4 bar and 235°C. The piston is pushed in by a force, F , until the cylinder volume is halved, and a cooling coil is used to maintain the pressure constant. If the dead state conditions are 1.013 bar and 22"C, determine:

(i) (ii) the change in availability; (iii) (iv) the irreversibility.

the work done on the gas by the force, F ;

the heat transfer from the gas;

Show the heat transfer process defined in part (iii) and the unavailable energy on a T-s diagram, and the work terms on a p-V diagram.

The following data may be used for helium, which can be assumed to behave as a perfect gas:

ratio of specific heats, K = cp/c, = 1.667 molecular weight, m, = 4 universal gas constant, 8 = 8.3143 kJ/ kmol K

[-3.154 kJ; 1.0967 kJ; -10.556 kJ; 2.0573 kJ]

5 A system at constant pressure consists of 10 kg of air at a temperature of 1000 K. Calculate the maximum amount of work that can be obtained from the system if the dead state temperature is 300 K, and the dead state pressure is equal to the pressure in the system. Take the specific heat at constant pressure of air, cp, as 0.98 kJ/kg K. (See Q.3, Chapter 1 .)

[3320.3 kJ]

6 A thermally isolated system at constant pressure consists of 10 kg of air at a temperature of 1OOOK and 10 kg of water at 3WK, connected together by a reversible heat engine. What is the maximum work that can be obtained from the system as the temperatures equalise? (See Q.4, Chapter 1.)

Assume

for water: c, = 4.2 kJ/kg K

c, = 0.7 kJ/kgK K = cp/c, = 1.0

for air: K = c,/c, = 1.4

[2884.2 kJ]

Page 62: Advanced Thermodynamics for Engineers

46 Availability and exergy

7 An amount of pure substance equal to 1 kmol undergoes an irreversible cycle. Neglecting the effects of electricity, magnetism and gravity state whether each of the following relationships is true or false, giving reasons for your assertion:

(i) f SQ = f (durn + P dvm)

> O (ii) dum + P dvm T

(iii) f 6W < f p dv,

where the suffix m indicates that the quantities are in molar terms. [False; false; true]

8 A gas turbine operates between an inlet pressure of 15 bar and an exhaust pressure of 1.2 bar. The inlet temperature to the turbine is 1500 K and the turbine has an isentropic efficiency of 90%. The surroundings are at a pressure of 1 bar and a temperature of 300 K. Calculate, for the turbine alone:

(i) the specific power output; (ii) the exhaust gas temperature; (iii) the exergy change in the gas passing through the turbine; (iv) the irreversibility or lost work.

Assume the working substance is an ideal gas with a specific heat at constant pressure of cp = 1.005 kJ/kg K and the specific gas constant R = 0.287 kJ/kg K.

[697.5 kJ/kg; 806K; 727.7 kJ/kg; 343.6 kJ/kg]

9 It is proposed to improve the energy utilisation of a steelworks by transferring the heat from the gases leaving the blast furnace at 600°C to those entering the furnace at 50°C (before the heat exchanger is fitted). The minimum temperature of the flue gases is limited to 150°C to avoid condensation of sulfurous acid in the pipework at exit pressure of 1 bar.

Draw a simple schematic diagram of the heat exchanger you would design, showing the hot and cold gas streams. Explain, with the aid of T-s diagrams, why a counterflow heat exchanger is the more efficient. If the minimum temperature differ- ence between the hot and cold streams is 1O"C, calculate the minimum loss of exergy for both types of heat exchanger, based on dead state conditions of 1 bar and 20°C.

[ -72.33 kJ/kg; -45.09 kJ/kg]

10 Find the maximum, and maximum useful, specific work (kJ/kg) that could be derived from combustion products that are (a) stationary, and (b) flowing, in an environment under the following conditions.

P(bar) TW) v(m3/kg) uW/kg) W / k g K)

Products (1) 7 1000 0.41 760.0 7.4 Environment (0) 1 298.15 0.83 289.0 6.7

[262.3; 220.3; 466.3; 424.31

Page 63: Advanced Thermodynamics for Engineers

3 Pinch Technology

In recent years a new technology for minimising the energy requirements of process plants has been developed: this has been named Pinch Technology or Process Integration by its major proponent, Linnhoff (Linnhoff and Senior, 1983; Linnhoff and Turner 1981). Fkcess plants, such as oil refineries or major chemical manufacturing plants, require that heating and cooling of the feed stock take place as the processes occur. Obviously it would be beneficial to use the energy from a stream which requires cooling to heat another which requires heating; in this way the energy that has to be supplied from a high temperature source (or utility) is reduced, and the energy that has to be rejected to a low temperature sink (or utility) is also minimised. Both of these external transfers incur a cost in running the plant. Pinch technology is an approach which provides a mechanism for automating the design process, and minimising the external heat transfers.

Pinch situations also occur in power generation plant; for example, in a combined cycle gas turbine (CCGT) plant (see Fig 3.1) energy has to be transferred from the gas turbine exhaust to the working fluid in the steam turbine. A T - s diagram of a CCGT plant is

Fig. 3.1 Schematic of CCGT

Page 64: Advanced Thermodynamics for Engineers

48 Pinch technology

shown in Fig 3.2, where the heat transfer region is shown: the pinch is the closest approach in temperature between the two lines. It is defined as the minimum temperature difference between the two streams for effective heat transfer, and is due to the difference in the properties of the working fluids during the heat transfer process (namely, the exhaust gas from the gas turbine cools down as a single phase but the water changes phase when it is heated) - this limits the amount of energy that can be taken from the hot fluid. The heat transfer processes are shown on a temperature-enthalpy transfer diagram in Fig 3.3, where the pinch is obvious.

Entropy, S

Fig. 3.2 T-s diagram of CCGT

fluid Pinch p i n t

/ Steam

I Q=AH I

Enthalpy

Fig. 3.3 H - T diagram of CCGT

Perhaps the easiest way of gaining an understanding of pinch techniques is to consider some simple examples.

Page 65: Advanced Thermodynamics for Engineers

A heat transfer network without a pinch problem 49

3.1 A heat transfer network without a pinch problem

This example has a total of seven streams, three hot and four cold, and it is required to use the heating and cooling potential of the streams to minimise the heat transfer from high temperature utilities, and the heat transfer to low temperature utilities. The parameters for the streams involved in the processes are given in Table 3.1. The supply temperature, T,, is the initial temperature of the stream, and the target temperature, TT, is the target final temperature that must be achieved by heat transfer. The heat flow capacity, mC, is the product of the mass flow and the specific heat of the particular stream, and the heat load is the amount of energy that is transferred to or from the streams.

Table 3.1 Specification of hot and cold streams

Supply Target temperature temperature Heat flow capacity Heat load

Stream no Stream type T,/(OC) TT/(OC) rnC/(MJ/hK) Q/(ryrJ/h)

1 Cold 95 205 2.88 2 Cold 40 220 2.88 3 Hot 310 205 4.28 4 Cold 150 205 7.43 5 Hot 245 95 2.84 6 Cold 65 140 4.72 7 Hot 280 65 2.38

316.8 518.4

-449.4 408.7

-426.0 354.0

-511.7

In this case there are three streams of fluid which require cooling (the hot streams) and four streams of fluid which require heating (the cold streams). The simplest way of achieving this is to cool the hot streams by transfemng heat directly to a cold water supply, and to heat the cold streams by means of a steam supply; this approach is shown in Fig 3.4. This means that the hot utility (the steam supply) has to supply 1597.9 MJ of energy, while the cold utility (a cold water supply) has to remove 1387.1 MJ of energy. Both of these utilities are a cost on the process plant. The steam has to be produced by burning a fuel, and use of the cold water will be charged by the water authority. In reality a minimum net heat supply of 1597.9-1387.1 = 210.8 MJ/h could achieve the same result, if it were possible to transfer all the energy available in the hot streams to the cold streams. This problem will now be analysed.

If heat is going to be transferred between the hot and cold streams there must be a temperature difference between the streams: assume in this case that the minimum temperature difference (dT,,,,,) is 10°C.

The method of tackling this problem proposed by Linnhoff and Turner (1981) is as follows.

Step I : Temperature intervals

Evaluate the temperature intervals defined by the 'interval boundary temperatures'. These can be defined in the following way: the unadjusted temperatures of the cold streams can be used, and the hot stream temperatures can be adjusted by subtracting ST,, from the actual values. In this way the effect of the minimum temperature difference has been included in the calculation. This results in Table 3.2.

Page 66: Advanced Thermodynamics for Engineers

50 Pinch technology

Hot streams $ $ $ Steam

I

9 u"

I50

354.0MJh 140 +

Cold water

205 95 65

Fig. 3.4 Direct heat transfer between the fluid streams and the hot and cold utilities

Table 3.2 Ordering of hot and cold streams

Supply Target temperature temperature

Stream no Stream type T,/(OC) TT/(OC) Adjusted temperatures Order ~~ ~~

1 Cold 95 95 T9 205 205 T5

220 220 T4

3 Hot 310 300 TI 205 195 T6

4 Cold 150 150 T7

5 Hot 245 235 T3 95 85 TlO

6 Cold 65 65 Tl1 140 140 T*

7 Hot 280 270 T2 65 55 T12

2 Cold 40 40 Tl 3

205 205 Duplicate

The parameters defining the streams can also be shown on a diagram of temperature against heat load (enthalpy transfer; see Fig 3.5). This diagram has been evaluated using the data in Table 3.2, and is based on the unadjusted temperatures. The hot stream line is based on the composite temperature-heat load data for the hot streams, and is evaluated using eqn (3.2); the cold stream line is evaluated by applying the same equation to the cold streams. It can readily be seen that the two lines are closest at the temperature axis, when they are still 25°C apart: this means that there is no 'pinch' in this example because the temperature difference at the pinch point is greater than the minimum value allowable.

Page 67: Advanced Thermodynamics for Engineers

400

350

300

250

200

150

100

50

0

~

~

Pinch

1 ' 1 I l l 1 I l l I I I l i

Fig. 3.6 Temperature intervals for heat transfer network

Page 68: Advanced Thermodynamics for Engineers

52 Pinch technology

Hence, the problem reduces to transferring energy from the hot streams to the cold streams, and finally adding 210.8 MJ/h from a hot utility. The mechanism for allocating the energy transfers will now be introduced.

Having defined the temperature intervals it is possible to consider the problem as shown in Fig 3.6. The energies flowing into and out of the combined systems, Qh and Qc, are those which have to be supplied by and lost to the external reservokc, respectively. It is also apparent that the difference between thcse values is the difference betwsen thz enthalpies of the hot and cold streams, i.e.

(3.1) Consideration wdl show that 6 H is constant, because the difference between the enthalpies

of the hot and cold streams is constant, and this means that any additional energy added from the hgh temperature supplies must be compensated by an equal amount of energy being rejected to the low temperature sinks: hence energy wdl have just flowed wastefully through the overall system. The heat transfer network can be shown schematically as in Fig 3.7. The heat flows through each of the temperature intervals can be evaluated as shown in the next step.

Q, - Q h = 6 H

Step 2: Interval heat balances

Table 3.2 includes the effect of the minimum temperature difference between the streams, ST,,,, and hence the intervals have been established so that full heat transfer is possible between the hot and cold streams. It is now necessary to apply the First Law to examine the enthalpy balance between the streams, when

where i = initial temperature of the interval i + 1 = final temperature of the interval

Applying this equation to this example results in the heat flows shown by the 6 H = mCdT values in Fig 3.8.

It can be seen from Fig 3.8 that the individual heat transfers are positive (Le. from the hot streams to the cold streams) in the first four intervals. In the fifth interval the amount of energy required by the cold streams exceeds that available from the hot streams in that temperature interval, but the energy can be supplied from that available in the higher temperature intervals. However, by the eighth interval the demands of the cold streams exceed the total energy available from the hot streams, and it is at this point that the energy should be added from the hot utility, because this will limit the temperature required in the hot utility. In reality, the 210.8 MJ could be provided from the hot utility at any temperature above 140"C, but the higher the temperature of the energy the more will be the irreversibility of the heat transfer process. It is now useful to look at the way in which the heat can be transferred between the hot and cold streams.

The streams available for the heat transfer processes are shown in Fig 3.9. First, it should be recognised that there is no heat transfer to the cold utility, and thus all the heat transfers from the hot streams must be to cold streams This constrains the problem to ensure that there is always a stream cold enough to receive heat from the hot sources. This means that the temperature of Stream 7 must be cooled to its target temperature of 65°C by transfemng heat to a colder stream: the only one available is Stream 2. Hence, the total heat transfer from Stream 7 is passed to Stream 2: an energy balance shows that the

Page 69: Advanced Thermodynamics for Engineers

A heat transfer network without a pinch problem 53

Hot reservoir (utilities)

Hot streams

SH = Q, - Q,

Cold streams Q,

Cold reservoir (utilities)

Fig. 3.7 Overall energy for heat transfer network

300 .1_1_1

4 0 5*

Streams mC6T CmC6T Q - .

3 128 4

3.7 233 1 128 4 Q -___ - - -

Q, -- 361 5 3,7,5 142 5

Q, I 504 0 3,7,5,2 99 3

-36 9 3,7,5, Q, - 603 3 --

2,1,4 Q - - 566 4

'75729 -358 7

207 8 -" 1,4

1

I ,6

6

Q6

7,5,2, -5 4 _I_-_

Q, 202 Heating 210 8 7,5,2, -236 8

7,5,2, -23 8 -344 1764 Q, ____--__ - -

I____ -582 1526 Q, ',6,2 -1044

QK -- -162 5 48 2 7,2 -5 0

2 -43 2

~ -167 5 43 2 Q,,

Q, --I I - -210 8 0

Fig. 3.8 Heat flows

temperature of Stream 2 is raised tc 218"C, and there is a residual heat capacity of 6.7 MJ/h before the stream reaches its target temperature In a similar manner it is necessary for Stream 5 to be matched to Stream 6 because this is the only stream cold enough to bring its temperature down to 95°C. In this way it is possible to remove Streams 6 and 7 from

Page 70: Advanced Thermodynamics for Engineers

54 Pinch technology

further consideration because they have achieved their target temperatures; Streams 2 and 5 must be left in the network because they still have residual energy before they achieve their targets. Figure 3.9 can be modified to Fig 3.10.

310 205 + E 95

2 220 0 b

65 b

.. ‘p 205 - r \ 0 4 U

mC

4 28

2 84

2 38

2 88

2 88

7 43

4 72

Q 449 4 MJh

426 0 M J h

511 7 M J h

316 8 MJh

51 8 4 M J h

408 7 M J h

354 0 MJ/h

Fig. 3.9 Initial heat transfer: heat transfer from Streams 5 to 6, and 7 to 2

310 300 205

245 220 b

205 4 I 4 220

mC

4 28

2 84

2 88

2 88

7 43

Q 449 4 M J h

72 0 MJ/h

316 8 M J h

6 7 M J h

408 7 MJ/h

Fig. 3.10 Removing the residuals from Streams 2 and 5

Streams 2 and 5 are represented in this diagram by their residual energies, and by the temperatures that were achieved in the previous processes. It is possible to cool Streams 3 and 5 by transfening energy with either of cold Streams 1 or 4. The decision in this case is arbitrary, and for this case Stream 3 will be matched with Stream 4, and Stream 5 will be matched with Stream 1. This results in the heat transfers shown in Fig 3.10, and by this stage Streams 5 and 4 can be removed from further consideration. This results in another modified diagram, Fig 3.1 1.

mC Q 4 2 8 4 0 7 MJ/h

2 8 8 2448 MJ/h

4 288 6 7 M J h 220 6 7 MJih 218

Cold streams

Fig. 3.11 Completing the heat transfer network

Page 71: Advanced Thermodynamics for Engineers

A heat transfer network without a pinch problem 55

By this stage it is necessary to consider adding the heat from the hot utility. In this case, the temperature at which this energy is added is relatively arbitrary, and the heat should be transferred at as low a temperature as possible. This is indicated by the 210.8 hU/h heat transferinFig 3.11.

The previous analysis has considered the problem in discrete parts, but it is now possible to combine all these sub-sections into a composite diagram, and this is shown in Fig 3.12.

MJ/h 140 U

4 28

2 84

2 38

95 2 8 8

40 2 88

150 7 4 3

65 4 7 2

449 4 MJih

426 0 MJih

511 7MJlh

3168MJih

518 4 MJih

408 7 M J h

354 0 MJ/h

Fig. 3.12 Composite diagram for heat exchanger network

The diagram given in Fig 3.12 suggests that seven heat exchangers are required, but the diagram is not the most succinct representation of the network problem, which is better illustrated as shown in Fig 3.13, which grows directly out of the original arrangement shown in Fig 3.4.

Hot streams

205 95 6 5

Fig. 3.13 Diagram showing the minimum number of heat exchangers to achieve heat transfer

Page 72: Advanced Thermodynamics for Engineers

56 Pinch technology

This case was a relatively straightforward example of a heat exchanger network in which it was always relatively easy to match the streams, because there was always a sufficient temperature difference to drive the heat transfer processes. The next example shows what happens when there is not sufficient temperature difference to drive the heat transfer processes.

3.2 A heat transfer network with a pinch point

Consider there are four streams of fluid with the characteristics given in Table 3.3. Assume that the minimum temperature difference between any streams to obtain acceptable heat transfer is 5"C, i.e. dT,, = 5°C. This value of dT,, is referred to as the pinch point because it is the closest that the temperatures of the streams are allowed to come. The temperatures of the streams can be ordered in a manner similar to the first case and the results of this are shown in Table 3.4.

Table 3.3 Characteristics of streams

Process stream Supply Target Heat capacity Heat load temperature temperature flowrate, mC mC(T, - TT)

Number Type T,/("C) TT/(OC) /(MJ/hK) /(MJ/h) ~~

1 Cold 50 110 2.0 120 2 Hot 130 70 3.0 180 3 Cold 80 115 4.0 140 4 Hot 120 55 1.5 97.5

Table 3.4 Definition of temperature intervals

Stream no Stream type T,/("C) TT/("C) Adjusted temperatures Order

1 Cold 50

2 Hot 130

3 Cold 80

4 Hot 120

110

70

115

55

50 T6 110 T3

125 TI 65 T5

80 T4 115 T2

115 Duplicate 50 Duplicate

These temperature intervals can be depicted graphically as shown in Fig 3.14. It can be seen that there are five intervals in this case, as opposed to the 12 in the first case.

It is now possible to draw the temperature-heat load diagram for this problem, and this is shown in Fig 3.15. It can be seen that the basic cold stream is too close to the hot stream and there will not be a sufficient temperature difference to drive the heat transfer processes. The modified cold stream line has been drawn after undertaking the following analysis, and does produce sufficient temperature difference at the pinch.

Page 73: Advanced Thermodynamics for Engineers

A heat transfer network with a pinch point 57

140

130

120

110

--. 100 0 v

90

8 80 b

k 70

60

50

40

Fig. 3.14 Temperature intervals for heat transfer network

-1 12.5 kJ/h

C o l d stream

-Basic cold s t reams

-Modified cold streams

I 1 1 , I

0 5 0 100 150 200 250 300

Heat load / (MJ/h)

Fig. 3.15 Temperature-heat load diagram, indicating pinch point

Step 3: Heat cascading

This step is an additional one to those introduced previously, and consideration will show why it comes about. Figure 3.16 indicates the heat flows in the various intervals and, in the left-hand column of figures, shows that if the heat flow from the hot utility (or heat source) is zero then there will be a negative heat flow of -12.5 units in temperature interval 3 between 110°C and 80°C. Consideration will show that this is impossible because it means that the heat will have been transferred against the temperature gradient. Such a situation can be avoided if sufficient energy is added to the system to make the largest negative heat flow zero in this interval. The result of this is shown in the right-hand column of figures in Fig 3.16, which has been achieved by adding 12.5 units of energy to

Page 74: Advanced Thermodynamics for Engineers

58 Pinch technology

the system. It can be seen that in both cases the difference between Q,, and Q, is 17.5 units of energy. If an energy balance is applied to the streams defined in Table 3.3 then

C(SH,=3x ( 1 3 0 - 7 0 ) + 1 . 5 ~ (120-55)-2~ ( l l 0 - 5 0 ) - 4 ~ (115-80)

= 17.5 (3.3) Hence, as stated previously, the energies obey the steady-flow energy equation for the system shown in Fig 3.16.

mC6T CmCGT -J!!Lpp{ + 12.5

42.5 6H = 30

30 -

6H = 2.5 32.5 - 45

31.5

30

Fig. 3.16 Temperature intervals and heat loading

There is now a point in the temperature range where the heat flow is zero: this point is called the pinch. In this example it is at 80"C, which means that the pinch occurs at a cold stream temperature of 80°C and a hot stream temperature of 85°C. There are three important constraints regarding the pinch:

Do not transfer heat across the pinch. Any heat flow across the pinch results in the same amount of heat being added to every heat flow throughout the system, and hence increases Q,, and Q,. Do not use the cold sink above the pinch. If the system has been designed for minimised heat flow it does not reject any heat above the pinch (see Fig 3.16, where the heat rejection has been made zero at the pinch). Do not use the hot source below the pinch. If the system has been designed for minimised heat flow it does not absorb any heat below the pinch.

It is hence possible to reduce the problem into two parts: above the pinch and below the pinch, as shown in Fig 3.17, which is a modification of Fig 3.7.

It is now necessary to break the problem at the pinch, and this results in Fig 3.18, which is the equivalent of Fig 3.9 for the first example.

Now the problem can be analysed, bearing in mind the restrictions imposed by the pinch point. This means that cooling to the utility stream is not allowable above the pinch, and hence the only transfer can be with the hot utility above the pinch. It is now necessary, as far as possible, to match the hot and cold streams above the pinch.

Consider Stream 2 is matched with Stream 1, and Stream 4 is matched with Stream 3. Then Stream 2 can transfer 70 MJ/h to Stream 1, and enable Stream 1 to achieve its

1.

2.

3.

1.

Page 75: Advanced Thermodynamics for Engineers

A heat transfer network with a pinch point 59

Hot streams

Cold streams

Cold reservoir (utilities)

Fig. 3.17 Breaking the problem at the pinch point

Q

135

52.5

70

140

Hot streams

n 130 I

70 _-----I 851 *

mC

3.0

1.5

2.0

4.0

Q

45

45

60

Cold streams

Fig. 3.18 Hot and cold streams with pinch

target temperature, while reducing its own temperature to 106.7"C. Also Stream 4 can transfer its energy to Stream 3 and this will raise the temperature of Stream 3 to 93.1"C. This shows that there is not sufficient energy available in these streams to achieve the target temperatures. The reason for this being an unsuitable approach is because the heat capacity flomte for Stream 1 above the pinch is greater than that of the cold stream above the pinch. Since both streams have the same temperature at the pinch point, then the higher temperature of the cold stream would have to be higher than that of the hot stream to achieve an energy balance: this would result in an impossible heat transfer situation. Hence, for a satisfactory result to be possible

mccold ~ m C , (3.4) above the pinch point. If the alternative match is used, namely, Stream 2 matched to Stream 3, and Stream 4 matched to Stream 1, then the answer shown in Fig 3.19 is obtained.

2.

Page 76: Advanced Thermodynamics for Engineers

60 Pinch technology

The heat transfers are shown on the diagram. It can be seen that it is not possible to match the hot and cold streams either above or below the pinch. This means that utility heat transfers are required from the hot utility above the pinch, and heat transfers to the cold utility are requeed below the pinch. This proposal does obey inequality (3.4), and is hence acceptable. It can be seen that, above the pinch point, energy has to be added to the system from the hot utility. This obeys the rules proposed above, and the total energy added is 12.5 MJ/h, which is in agreement with the value calculated in Fig 3.16. Considering the heat transfers below the pinch, it can be seen that Stream 1 can be heated by energy interchange with Streams 2 and 4: neither Stream has sufficient capacity alone to bring Stream 1 to its target temperature. However, it is feasible to bring about the heating because

3.

4.

mChot 2 mCcold (3.5) which is the equivalent of inequality (3.4) for the transfers below the pinch. In this case it was chosen to transfer all the energy in Stream 4 because this results in a lower temperature for heat transfer to the cold utility. The 30 MJ/h transferred to the cold utility is in line with that calculated in Fig 3.16.

These diagrams can now be joined together to give the composite diagram in Fig 3.19.

mC Q Hot streams

n n 3 .O 180 70 ‘3O

1.5 97.5

106.3 2.0 130 U

1.5 M l h 52 .5MJh 45MJh ISMJlh

4.0 140 5 M l h 135MJh

Cold streams

Fig. 3.19 Composite diagram for heat transfer network in Example 2

The heat load against temperature diagram for this problem, before heat transfer from the utilities has been supplied, is shown in Fig 3.15, and was discussed previously. It is now possible to consider the modified diagram, when it can be seen that the energy transfers have produced a sufficient temperature difference to satisfy the constraints of the problem.

3.3 Concluding remarks

A method has been introduced for improving the efficiency of energy transfers in complex plant. It has been shown that in some plants there is a pinch point which restricts the freedom to aansfer energy between process streams. To ensure that the plant attains its maximum efficiency of energy utilisation, energy should be added to the system only above the pinch, and extracted from it only below the pinch.

Page 77: Advanced Thermodynamics for Engineers

Problems 61

PROBLEMS

1 A process plant has two streams of hot fluid and two streams of cold fluid, as defined in Table P3.1. It is required to minimise the energy which must be transferred to hot and cold utilities by transferring energy between the streams. If the minimum temperature difference for effective heat transfer is 20"C, design a network which achieves the requirement, and minimises the transfers to the utilities. Is there a pinch point in this problem, and at what temperature does it occur? Calculate the minimum heat transfers to and from the cold and hot utilities.

Table P3.1 Data related to Q.l

Supply Target temperature temperature Heat flow capacity

Stream no Stream type T,/("C) T,/("C) rnC/(MJ/hK)

1 Hot 205 65 2.0 2 Hot 175 75 4.0 3 Cold 45 180 3.0 4 Cold 105 155 4.5

[105"C (cold stream); e",,. = 90 M / h ; Qc,,, = 140 M / h l

2 Some stream data have been collected from a process plant, and these are listed in Table P3.2. Assuming the minimum temperature difference between streams, ATrnh = 1O"C, (a) calculate the data missing from Table P3.2; (b) analyse this data to determine the minimum heat supplied from the hot ytility, the

(c) draw a schematic diagram of the heat transfer network.

Data related to Q.2

minimum heat transferred to the cold utility, and the pinch temperatures;

Table P3.2

Supply Target temperature temperature Enthalpy change Heat flow capacity

rnC/(kW/K) Stream no Stream type T,/("C) T,/("C) AH/(kW) ~ ~~

1 Cold 60 180 ? 3 2 Hot 180 40 ? 2 3 Cold 30 130 220 ? 4 Hot 150 40 400 ? 5 Cold 60 80 40 ?

[360 kW; 280 kW; 2.2 kW/K; 4.0 kW/K; 2.0 kW/K; e",,, = 60 kW; e,,,, = 160 kW; TCp,nch = 140"C]

3 Figure P3.3 shows a network design using steam, cooling water and some heat recovery. (a) Does this design achieve the minimum energy target for AT,,, = 20"C? (b) If the current network does not achieve the targets, show a network design that

[(a) TCplnch = 150°C; THplnch = 170OC; Q, = 480 kW; QH = 380 kW; (b) Q,,,, = 360 kW; does.

QH,,, = 260 k w 1

Page 78: Advanced Thermodynamics for Engineers

62 Pinch technology

Cold water duty: 480 kW Cold water duty: 480 kW

260" - 260" 60"

420 kW 300 kW

Process

L I

Fig. P3.3 Network for Q.3

60"

420 kW 300 kW

4 Figure P3.4 shows two hot streams and two cold streams for heat integration (subject to AT,, = 20°C).

(i) (ii)

What are the energy targets? Show a network design achieving these targets.

[QH,, = 0; Qc,,,, = 01

Process

mC = 20 m C = 10

t '55"c t

50"

130°C

r!l

150"

30c

&I

50"

Cold streams Fig. P3.4 Network for 4.4

150"

5 Figure P3.5 shows an existing design of a process plant, containing two exothermic processes. These require streams of reactants as shown in the diagram, and produce products at the temperatures shown. The plant achieves the necessary conditions by providing 480 k W of heat from a steam source, and rejects a total of 560 k W of energy to cold water utilities; only 460 k W is transferred between the streams.

(a) Show that there is a pinch point, and evaluate the temperature. (b) Show that the existing plant is inefficient in its use of the energy available. (c) Calculate the energy targets for AT^,, = 20°C and show a design that achieves these

targets.

TCpinch = 1 10°C; THplnch = 130°C; (b) Q, = 560 kW; QH = 480 kW, (c) QCmm = 210 kW; QH,, = 130 kW]

Page 79: Advanced Thermodynamics for Engineers

Problems 63

A H = 4 6 0 k W AH = 360 kW

Fig. P3.5 Network for Q.5

steam 60 kW

70" /7f 150" 180" /7( 30"

6 Recalculate the problem in Q.5 using a AT,,,, = 10°C. Comment on the effect of reducing the minimum temperature difference.

[(a) TCpmch = 110°C; THpmch = 120°C; (b) Q, = 560 kW; Q H = 480 kW; (c) Q,,,, = 120 kw; eHmln = 40 kw1

Process * Process

7 A network for a process plant is shown in Fig P3.7.

(a) Calculate the energy targets for ATrnn = 10°C and show a design that achieves these

(b) Explain why the existing network does not achieve the energy targets. targets.

[(a) QCmln = 190 kW; QH,,T= 90 kW; (b) there is transfer across the pinch]

- 190"

660 kW 420 kW

110" 160"

Fig. P3.7 Network for Q.7

Page 80: Advanced Thermodynamics for Engineers

Rational Efficiency of a Powerplant

4.1 The influence of fuel properties on thermal efficiency The thermal efficiency of a cycle has been defined previously, in terms of specific quantities, as

w,t 9 t h = -

qin

where wnet = net work output from the cycle per unit mass of fluid, and qin = energy addition to the cycle per unit mass of fluid.

In this case qin is the energy transfer to the working fluid, and does not take into account any losses in the boiler or heat transfer device. Equation (4.1) can be rewritten for the whole powerplant, including the boiler or heat transfer mechanism, as

(4.2) W",t

% ' q B V t h = - - Ah0

where l;lo = overall efficiency of powerplant, l;lB = efficiency of boiler, qth = thermal efficiency of cycle,

wWt = net work output from the cycle per unit mass of fuel, and Ah,, = specific enthalpy of reaction of fuel.

This might be considered to be an unfair, and possibly misleading, method of defining the efficiency because the energy addition cannot all be turned into work, as was shown when considering exergy and availability. Another definition of efficiency can be derived based on the Second Law, and this relates the work output from the cycle to the maximum work output obtainable.

The efficiency of the powerplant has been related, in eqn (4.2), to the amount of energy that has been added to the cycle by the combustion of the fuel. In the past this has been based on the enthalpy of reaction of the fuel, or usually its calorific value, Qi. It was shown previously (Chapter 2) that this is not the energy available for the production of work, and that the maximum available work that can be obtained from the fuel is based on

Page 81: Advanced Thermodynamics for Engineers

Rational efficiency 65

the change of its exergy at the dead state conditions. Hence, the maximum available work from unit mass of fuel is

-Ago = g, - g,

Ago = Ah0 - To(s, - SPo>

(4.3)

This is related to the enthalpy of formation by the equation

(4.4)

It was shown in Chapter 2 that I Ago I could be greater than or less than I Aho 1 , and the difference was dependent on the structure of the fuel and the composition of the exhaust products. The efficiency of the powerplant can then be redefined as

(4.5) Wnet a=-

where wmt = actual net work output from the cycle per unit mass of fuel,

- Ago

and Ago = change of Gibbs function caused by combustion, = maximum net work obtainable from unit mass of fuel.

Equation (4.5) is often referred to as the Second Law Efficiency, because the work output is related to the available energy in the fuel, rather than its enthalpy change. The actual effect on thermal efficiency of using the change of Gibbs function instead of the enthalpy of reaction is usually small (a few per cent).

4.2 Rational efficiency

When the efficiencies defined in eqns (4.2) and (4.5) are evaluated they contain terms which relate to the 'efficiency' of the energy transfer device (boiler) in transferring energy from the combustion gases to the working fluid. These effects are usually neglected when considering cycles, and the energy added is related to the change in enthalpy of the working fluid as it passes through the boiler, superheater, etc. Actual engine cycles will'be considered later. First, a general heat engine will be considered (see Fig 4.1). For convenience the values will all be taken as specific values per unit mass flow of working fluid.

The engine shown in Fig 4.1 could be either a wholly reversible (i.e. internally and externally) one or an irreversible one. If it were internally reversible then it would follow the Camot cycle 1-2-3s-4-1. If it were irreversible then it would follow the cycle 1-2-3- 4- 1. Consider first the reversible cycle. The efficiency of this cycle is

which is the efficiency of an internally reversible engine operating between the temperature limits TL and TL. This efficiency will always be less than unity unless TL = 0. However, the Second Law states that it is never possible to convert the full energy content of the energy supplied into work, and the maximum net work that can be achieved is

(4.7) G m t = b 2 - b , = h2- hi - TO(s2- $1)

where To is the temperature of the dead state.

Page 82: Advanced Thermodynamics for Engineers

66 Rational efficiency of a powerplant

Fig. 4.1 A heat engine operating between two reservoirs: (a) schematic diagram of the engine; (b) T-s diagram for the engine which receives and rejects energy isothermally; (c) T-s diagram for

the engine which receives and rejects energy non-isothermally

Hence, the Second Law efficiency of the heat engine is

where q2,* is called the Second Law efficiency, or the rational efficiency, qR. This result can be interpreted in Fig 4.1 (b) in the following way. Examination of eqn (4.8) shows that if the T3s = To then l;lR = 1.0. This because when T3s = To then area C is zero (Le. the line at TL = To is coincident with that at T3s). Then the difference in the enthalpies, h2 - h, , is depicted by areas A + B, and the unavailable energy To (s2, s l ) by area B, and hence the energy available to produce work is area A. This shows that the cycle is as efficient as an internally reversible cycle operating between the same two temperature limits (T3s and Tl). If To is not equal to T3, but is equal to a lower temperature TL, then the rational efficiency will be less than unity because energy which has the capacity to do work is being rejected. This is depicted by area C in Fig 4.l(b), and the rational

Page 83: Advanced Thermodynamics for Engineers

Rational efJiciency 67

efficiency becomes

Area A

Areas (A + C) < 1 - - Areas(A + B + C) - Areas(B + C)

Areas (A + B + C) - Areas (B) ?R = (4.9)

Consideration of eqn (4.8) shows that it is made up of a number of different components which can be categorised as available energy and unavailable energy. This is similar to exergy, as shown below:

b = a - a - h - h , - To ( s - so ) = E - Q, O--- -+ Y ",,av&,ble available unavailable

energy energy energy ene%Y

(4.10)

(4.13)

If To = T3s then the rational efficiency, qR = 1.0. If To < T3,, as would be the case if there were external irreversibilities, then vRc 1.0 because the working fluid leaving the engine still contains the capacity to do work.

If the engine is not internally reversible then the T-s diagram becomes 1-2-3-4- 1, as depicted in Fig 4.1 (b). The effect of the internal irreversibility is to cause the entropy at 3 to be bigger than that at 3s and hence the entropy difference 6S,> 6 S I 2 . The effect of this is that the net work becomes

w, = ( h 2 - h,) - (h3 - h4) = T I ( S 2 - s1) - T3b3 - s4)

= - SI) - T3(s3s - $4) - T303 - s 3 s )

= h 2 - f i l - T o ( s 2 - $1) - v 3 - T o 1 0 2 - $ 1 ) - T 3 0 3 - s 3 s )

= 6 4 2 - 6Q,P,, - ( T 3 - T o ) ( s 2 - $ 1 ) - T3(s3 - s 3 s ) (4.14) Hence, from eqn (4.14)

(4.15)

In the case of the irreversible engine, even if To = T3 the rational efficiency is still less than unity because of the increase in entropy caused by the irreversible expansion from 2 to 3. The loss of available energy, in this case the irreversibility, is depicted by area D in Fig 4.1 (b). In the case of To = T3,

(4.16)

If the dead state temperature is less than T3 then the rational efficiency is even lower because of the loss of available energy shown as area C in Fig 4.1 (b).

Page 84: Advanced Thermodynamics for Engineers

68 Rational eficiency of a powerplant

UI; until now it has been assumed that the cycle is similar to a Camot cycle, with isothermal heat supply and rejection. Such a cycle is typical of one in which the working fluid is a vapour which can change phase. However, many cycles use air as a working fluid (e.g. Otto, Diesel, Joule cycles, etc), and in this case it is not possible to supply and reject heat at constant temperature. A general cycle of this type is shown in Fig 4.1 (c), and it can be seen that the heat is supplied over a range of temperatures from TI to T2, and rejected over a range of temperatures from T3 to T4. If only the heat engine is considered then it is possible to neglect the temperature difference of the heat supply: this is an external irreversibility. Furthermore, to simplify the analysis it will be assumed that TL = To = T4.

However, it is not possible to neglect the varying temperature of heat rejection, because the engine is rejecting available energy to the surroundings. If the cycle is reversible, i.e. 1-2-3s-4- 1, then the rational efficiency of the cycle is

(4.17) Wnet V R = ~

b2 - bl

If the cycle shown is a Joule (gas turbine) cycle then

Wnet = h2 - hI - (h3s - h4) = b2 - b l + - $1) - { b3s - b4 + - s 4 ) 1 = b, - b, - (b3s - b4) (4.18)

Hence, the net work is made up of the maximum net work supplied and the maximum net work rejected, i.e.

wnet = { *netlsupplie~ - [*netlrcjected (4.19)

These terms are defined by areas 1-2-3s-5-4-1 and 4-3s-5-4 respectively in Fig 4.1 (c). Thus the rational efficiency of an engine operating on a Joule cycle is

(4.20)

Equation (4.20) shows that it is never possible for an engine operating on a cycle in which the temperature of energy rejection varies to achieve a rational efficiency of 100%. This is simply because there will always be energy available to produce work in the rejected heat.

If the cycle is not reversible, e.g. if the expansion is irreversible, then the cycle is defined by 1-2-3-4-1, and there is an increase in entropy from 2 to 3. Equation (4.18) then becomes

wnet = h2 - hl - (h3 - h4) = b2 - bl + To(s2 - 31) - {b? - b4 + TO(s3 - sq)}

= b2 - b, - (b3 - b4) - To($, - sg,) (4.21) --- area area area

1-2-3 s - 5 4 1 6-3- 5 4 7 - 8 - 5

Equation (4.21) shows that the irreversibility of the expansion process reduces the net work significantly by (i) increasing the amount of exergy rejected, and (ii) increasing the irreversibility of the cycle. The rational efficiency of this cycle is

Page 85: Advanced Thermodynamics for Engineers

Rankine cycle 69

The irreversible cycle can be seen to be less efficient than the reversible one by comparing eqns (4.20) and (4.22). In the case shown b2 - b, is the same for both cycles, but b3 - b4 > b3, - b4, and in addition the irreversibility To(~3 - s3J has been introduced.

4.3 Rankine cycle

It was stated above that the efficiency of a cycle is often evaluated neglecting the irreversibility of the heat transfer to the system. Such a situation can be seen on the steam plant shown in Fig 4.2(a), which can be represented by the simplified diagram shown in Fig 4.2(b), which shows the heat engine contained in system B of Fig 4.2(a).

Now, consider the passage of the working fluid through system A. It enters system A with a state defined by 3 on the Rankine cycle (see Fig 4.3), and leaves system A with a state defined by 2. The usual definition of thermal efficiency given in eqn (4.1) results in

(4.23) w,t Vth = -

h3 - h2

where w , ~ = specific net work output from the cycle.

that is obtainable from the fluid is However, from the definition of exergy, the maximum work (per unit mass of steam)

v C , ~ = b3 - b2 = (h3 - h2) - To(~3 - $ 2 ) (4.7)

Examination of eqn (4.7) shows that it consists of two terms: a difference of the enthalpies between states 3 and 2, and the product of the dead state iemperature, To, and the difference of the entropies between states 3 and 2. The first term is obviously equal to the energy added between states 2 and 3, and is q,,. Considering the other term, the energy rejected to the cold reservoir is

4 0 U t = h4 - h I = TI ( $ 4 - SI 1 (4.24)

If the condenser were of infinite size then the temperature at 1 could be equal to the environmental temperature, To. Also, if the feed pump and the turbine were both reversible then processes 3-4 and 1-2 would be isentropic and eqn (4.24) could be reduced to

qout = h4 - h , = To(s3 - $ 2 ) (4.25)

Hence, for an internally reversible engine the maximum net work output is the sum of the heat supplied and the heat rejected, as shown previously simply from considerations of energy. If the cycle was not reversible due to inefficiencies in the turbine and feed pump, then the difference between the entropies at states 4 and 1 would be greater than the differences at states 2 and 3, and qoUt would be greater than in the ideal case. Hence, the cycle would be less efficient.

The thermal efficiency of an ideal Rankine cycle (i.e. one in which the turbine and feed pump are both isentropic) is given by

(4.26)

Page 86: Advanced Thermodynamics for Engineers

70 Rational efficiency of a powerplant

w

f i5 F

(a) (b) Fig. 4.2 Steam turbine power plant

3

i

W T \

Qi.

i i I i I 'I'..;L i - \ \., 1,

\

/ = \\,\,

'\ 2 4

Page 87: Advanced Thermodynamics for Engineers

Examples 71

This means that a steam turbine which operates on a reversible Rankine cycle will have a rational efficiency of 100%. If there are any irreversibilities then the rational efficiency will be less than 100%. Rational efficiency shows the scope for improving the device within the constraints of, say, peak pressure and temperature. If the rational efficiency is low then the efficiency can be improved significantly, whereas if it is high then not much improvement is possible.

It has to be remembered that these definitions of rational efficiency for a steam plant cycle are based on the dead state temperature being made equal to the condenser temperature. If another temperature is chosen then the rational efficiency will be reduced as shown in eqn (4.22). It will be shown that this results in a rational efficiency of unity for an ideal (internally reversible) Rankine cycle, and such an approach can be used to indicate how far an actual (irreversible) steam plant cycle falls short of the yardstick set by the reversible one. However, this approach also masks the 'cost' of external irreversibilities between the working fluid in the condenser and the true dead state of environmental conditions. These effects are discussed in the examples.

When the term rational efficiency is applied to an air-standard cycle it is never possible to achieve a value of unity because the temperature at which energy is rejected is not constant. This will also be considered in the examples by reference to the Joule cycle for a gas turbine.

4.4 Examples

Q.1 Steam turbine cycles A steam turbine operates on a basic Rankine cycle with a maximum pressure of 20 bar and a condenser pressure of 0.5 bar. Evaluate the thermal efficiency of the plant. Calculate the maximum net work available from the cycle, and evaluate the rational efficiency of the cycle.

Solution

The T-s diagram for the Rankine cycle is shown in Fig 4.4. The parameters for the state points on cycle 1-2-3-4-6- 1 will now be evaluated.

Conditions at 4

p4=20 bar; X = 1 t, = 212.4"C; h, = 2799 kJ/kg; sg = 6.340 kJ/kg K

Conditions at 6 Pa= 0.5 bar; s6 = 6.340 kJ/kg K t, = 81.3"C; sg = 7.593 kJ/kg K;

h, = 2645 kJ/kg; h f = 340 kJ/kg sf = 1.091 kJ/kg K

Thus

6.340 - 1.091 7.593 - 1.091

X g = = 0.8073

h, = xh, + (1 - x)h, = 0.8073 x 2645 + ( 1 - 0.8073) x 340 = 2200.8 kJ/kg

Page 88: Advanced Thermodynamics for Engineers

72 Rational eficiency of a powerplant

k-

5- E

F ?

To, =T,=8 I .3 "C

T,,=20°C saturated vapotn

.]me '.. --_ ~~

Entropy, S

Fig. 4.4 Temperature-entropy diagram for a basic Rankine cycle

Conditions at 1

p 1 = 0.5 bar; s1 = 1.091 kJ/kg K; h, = 340 kJ/kg v1 = v f = 0.001029 m3/kg

Conditions at 2

p 2 = 20 bar; S, = 1.091 kJ/kg K v2 = 0.001 157 + 0.2 x (-0.0009 x = 0.001 155 m3/kg

Hence feed pump work,

wp= w12= - v d p = -dhlz = - (0.001029 + 0.001 155)/2 x 19.5 x 100 = -2.1 1 kJ/kg

h2=h,+dh12=340+2.11 =342.1 kJ/kg

Work done by turbine,

W T = h4 - h, = 2799 - 2200.8 = 598.2 kJ/kg K

Thermal efficiency of the cycle is

w,,~ wT + wP 598.2 - 2.1 1

qh h4 - hZ 2799 - 342.1 - = 0.243 Vth = __. = ~ -

The maximum net work available can be evaluated from the change of exergy between state points 4 and 2. Hence

Gne1 = b, - b,

It is necessary to define a dead state condition. This is arbitrary, and in this case will be taken as the condition at 1. Hence, p , = 0.5 bar, To = 81.3 + 273 = 354.3 K. Thus

b4 = h4 - T0s4 - ( h , - Toso) = 2799 - 354.3 x 6.340 - a,-, = 552.7 - a, kJ/kg b, = h, - Tos2 - (h, - Toso) = 342.1 - 354.3 x 1.091 - a, = -44.4 - a, kJ/kg

Page 89: Advanced Thermodynamics for Engineers

Examples 73

Hence Gmt = 552.7 - (-44.4) = 597.1 W/kg

Thus, the rational efficiency is wnet wnet 598.2 - 2.1 1

R - = 1.00 77 --=-= b4-b2 %et 597.1

In this case the rational efficiency is equal to unity because the turbine and feed pump have isentropic efficiencies of loo%, and it was assumed that the temperature of the working fluid in the condenser was equal to the dead state (ambient) temperature. Hence, although the cycle is not very efficient, at 24.3%, there is no scope for improving it unless the operating conditions are changed.

It is possible to evaluate the rational efficiency of a steam plant operating on a Rankine cycle in which the condenser temperature is above the ambient temperature. If the dead state temperature in the previous example was taken as 20°C rather than 81.3"C, then the following values would be obtained:

b, = h, - Tos4 - a, = 2799 - 293 x 6.340 - a, = 941.4 - a, kJ/kg b2 = h2 - To$, - a, = 342.1 - 293 x 1.091 - U , = 22.4 - a, W/kg

Hence Gnet = 941.4 - 22.4 = 919.0 W/kg

Thus, the rational efficiency is

wnet wnet 598.2 - 2.1 1 l ; l R = - - - = 0.649

b4-b2 knet 919.1 This result shows that irreversibilities in the condenser producing a temperature drop of

81.3"C to 20°C would reduce the potential efficiency of the powerplant significantly. Basically this irreversibility is equivalent to a loss of potential work equal to the area of the T-s diagram bounded by the initial dead state temperature of 81.3"C and the final one of 20°C and the entropy difference, as shown in Fig 4.4, i.e. (To, - T,)(s4 - s,), which is equal to (354.3-293) x (6.340 - 1.09) = 321.8 W/kg.

Q.2 Steam turbine cycles Re-evaluate the parameters for the steam plant in Q.l based on To = 81.3"C if (a) the isentropic efficiency of the turbine is 80%; (b) the isentropic efficiency of the feed pump is 70%; (c) the efficiency of the components is the combination of those given in (a) and (b).

Solution

(a) Turbine efficiency, l;lT = 80%. If the turbine efficiency is 80% then the work output of the turbine becomes

W T = 77T(wT)isen = 0.80 x 598.2 = 478.6 H/kg Hence, the thermal efficiency of the cycle is

wXt W T + wp 478.6 - 2.1 1 qh h4 - h, 2799 - 342.1

t h - - = 0.194 77 --=--

Page 90: Advanced Thermodynamics for Engineers

74 Rational eficiency of a powerplant

The rational efficiency is

[ ~ n c t l a c t u a ~ - [wnetlactual - 478-6-2-11 V R = - - = 0.80

b4 - b2 %et 597.1

The rational efficiency has been significantly reduced by the inefficiency of the turbine, and the rational efficiency is approximately equal to the isentropic efficiency of the turbine.

(b) Feed pump efficiency, qp = 70%. The feed pump work becomes

h,, = h, + dh,,. = 340 + 3.01 = 343 kJ/kg

The thermal efficiency of the cycle becomes

wmt W T + wp 598.2 - 3.01 qi,, h4-hZ 2799-343

rh - = 0.242 V =-=--

The exergy at 2' may be evaluated approximately by assuming that the entropy does not change significantly over the pumping process, i.e. s2' = s,. Hence

b,. = h2( - T0s2. - a, = 343 - 354.3 x 1.091 - a, = -43.5 - a, H/kg

The rational efficiency is

= 0.998 [Wnetlactual - [Wnetlactual - 598.2-3.01 - - VR = b4 - bZ k t 596.2

The reduction in thermal efficiency is small, and the rational efficiency is almost 1. Hence, the Rankine cycle is not much affected by inefficiencies in the feed pump. Turbine efficiency l;lT = SO%, feed pump efficiency qp = 70%. The thermal efficiency is

(c)

wnet WT + wp 478.6 - 3.01 qi,, h4-hT 2799-343

v = - = - - rh - = 0.194

The rational efficiency is

Q.3 Steam turbine cycles

The steam plant in Q.2 is modified so that the steam is superheated before entering the turbine so that the exit conditions from the turbine of the ideal cycle are dry saturated. Evaluate the parameters for the steam plant if (a) the isentropic efficiency of the turbine is 80%; (b) the isentropic efficiency of the feed pump is 70%; (c) the efficiency of the components is the combination of those given in (a) and (b). Assume that to = 81.3"C. This cycle is shown in Fig 4.5.

Page 91: Advanced Thermodynamics for Engineers

Examples 75

Fig. 4.5 Temperature-entropy diagram for a Rankine cycle with superheat

Conditions at 6

p6 = 0.5 bar t, = 81.3"C; h, = 2645 kJ/ kg; sg = 7.593 kJ/kg K; h6 = 2645 kJ/kg;

hf = 340 kJ/kg sf = 1.091 kJ/kg K

s6 = 7.593 H/kg K Conditions at 5

p5 = 20 bar; s5 = s6 = 7.593 H/kg K

Hence

7.593 - 7.431

7.701 - 7.431 t , = 5 0 0 + x 100 = 560°C

h5 = 3467 + 0.6 x (3690 - 3467) = 3601 W/kg

Work done by turbine,

WT = h5 - h6 = 3601 - 2645 = 956 kJ/kg K Thermal efficiency of the cycle is

w , ~ w T + w ~ 956-2.11 vth=-=-= = 0.292

qh h5-h2 3601 -342.1

b5 = h5 - TOSS - U O = 3601 - 354.3 x 7.593 - a0 = 910.8 - a, kJ/kg b, = h, - TOsz - a0 = 342.1 - 354.3 x 1.091 - a0 = -44.4 - a0 kJ/kg

The rational efficiency is

= 1.00 [~net lac tua l - [wnetlactua~ - 956-2-11 - - 955.2

v R = bq - b2 %et

Page 92: Advanced Thermodynamics for Engineers

76 Rational eficiency of a powerplant

(a) Turbine efficiency, l;lT = 80%.

WT = vT(wT)isen = 0.80 x 956 = 764.8 H/kg

Hence, the thermal efficiency of the cycle is

wnet WT + wp 764.8 - 2.1 1

qi. hq - h2 3601 - 342.1 fh - = 0.234 77 =-=--

The rational efficiency is

= 0.799 [ ~ n e t l a c t u a l - netl lac ma^ - 764-8-2-11 - - 77R = bs - b2 +net 955.2

Hence, again, the effect of the turbine efficiency is to reduce the rational efficiency by an amount almost equal to its isentropic efficiency.

(b) Feed pump efficiency, l;lp = 70%. The effect on the feed pump work is the same as above, and the thermal efficiency of the cycle becomes

w,, w T + w ~ 956-3.01 qi. h, - h2' 3601 - 343

ih - = 0.295 77 =-=--

The rational efficiency is

956 - 3.01 910.8 - (-44.4)

= 0.998 [Wnetlactual - [wnetlactual - - - 77R = bs - b2,

Again, the reduction in thermal efficiency is small, and the rational efficiency is almost 1.

(c) Turbine efficiency, ~ ? , = 8 0 % ; feed pump efficiency, 7;lp=70%. The thermal efficiency is

wne1 WT + wp 764.8 - 3.01 VIh=-=-- - = 0.234

qi. h5 - h2, 3601 - 343

The rational efficiency is

[WnetIactua~ - [WnetIactual - 764.8 - 3.01 = o.798 b5 - b2' %net

- - 955.2

It can be seen that both the basic and superheated Rankine cycles are equally affected by inefficiencies in the individual components. The rational efficiency is an estimate of how close the cycle comes to the internally reversible cycle.

77R =

Q.4 Gas turbine cycle

A gas turbine operating on an ideal Joule cycle has a pressure ratio of 20: 1 and a peak temperature of 1200 K. Calculate the net work output, the maximum work output, the thermal efficiency, and the rational efficiency of the cycle. Assume that the working fluid is air with a value of K = 1.4 and a specific gas constant R = 0.287 H/kg K. The inlet conditions at 1 are 1 bar and 300 K, and these should be taken as the dead state conditions also.

Page 93: Advanced Thermodynamics for Engineers

Examples 77

Entropy, S

Entropy, S @)

Fig. 4.6 (a) Temperature-entropy diagram for Joule cycle. (b) Temperature-entropy diagram for Joule cycle with irreversible compression and expansion

Solution

The ideal Joule cycle is depicted by 1-2s-3-4s-1 in Fig 4.6(a). The relationship between entropy and the primitive properties for an ideal gas is

T P TO Po

s - % = cPln - - R In-

From the parameters given,

KR 1.4 x 0.287 - = 1.0045kJ/kgK. c p = - - 0.4 K-1

Page 94: Advanced Thermodynamics for Engineers

78 Rational efficiency of a powerplant

The compression process from 1 to 2s is isentropic, and hence

The isentropic work done in the compressor is

wq,,= -dhl,= cp(T1- TZS) = 1.0045 x (706.1 - 300) ~407.9 kJ/kg The energy added to the cycle is

q Z s 3 = cP( T3 - TZs) = 1.0045 x (1 200 - 706.1) = 496.1 kg For an isentropic expansion from 3 to 4s

(K - 1 ) / K 0.4/1.4

T4s = T 3 ( 9 = 1200 x ($) = 509.9K

Hence the isentropic turbine work is

wT,,, = cP (T3 - TaS) = 1.0045 x (1200 - 509.9) = 693.2 kJ/kg The net work from the cycle is wmt = wT + wc = 693.2 + (-407.9) = 285.3 kJ/kg, and the thermal efficiency is

This value is equal to the standard expression for the efficiency of a Joule cycle,

The maximum net work output is defined by

$,,,.t = b3 - b2 Now, the values of exergy are defined, for a perfect gas, as

b = ( h - T ~ s ) - (ho - TOSO) = (h - ho) - T ~ ( s - S O )

= cp { (T - To) - To ( ln- - - “ h q } To K Po

Thus, along an isobar

( T - T o ) - T o h - TO

Since both 2s and 3 are at the same pressure

Thus

- 706.1) - 300h 706.1

Page 95: Advanced Thermodynamics for Engineers

Examples 79

The rational efficiency of the cycle based on these dead state conditions is

wnet 285.3

Gnet 336.3 9R=-=-- - 0.848

This means that the ideal Joule cycle is only capable of extracting 84.8% of the maximum net work from the working fluid, whereas under similar conditions (with the dead state temperature defined as the minimum cycle temperature) the Rankine cycle had a rational efficiency of 100%. The reason for this is that the energy rejected by the Joule cycle, from 4s to 1, still has the potential to do work. The maximum net work obtainable from the rejected energy is b,, - b, . This is equal to

509.9 = 1.0045 x (509.9 - 300) - 300111 - 1 300

This term is shown in Fig 4.6(a), and is energy which is uaavailable for the production of work.

If the efficiencies of the turbine and compressor are not 100% then the diagram is shown in Fig 4.6(b). It can be seen that the rejected energy is larger in this case.

Q.5 Gas turbine cycle

For the gas turbine cycle defined in Q.4, calculate the effect of (a) a turbine isentropic efficiency of 80%; (b) a compressor isentropic efficiency of 80%; and (c) the combined effect of both inefficiencies.

Solution

(a) Turbine isentropic efficiency, l;lT = 80%. This will affect the work output of the turbine in the following way:

W T = ~ T W X , , = 0.8 x 693.2 = 554.6 W/kg

Hence

w,,~ = WT + W C = 554.6 + (-407.9) = 146.7 kJ/kg

and

wnet 146.7 %I=-=- = 0.296

q 2 s 3 496.1

The temperature after the turbine is T4 = T3 - vTATsn = 1200 - 0.8 x 690.1 = 647.9 K, and the entropy at 4, related to 1, is

647.9 - 0.7734 kJ/kg K s4 - s1 = cpln- - R ln- = 1.0045 In - - T4 P4

TI PI 300

Page 96: Advanced Thermodynamics for Engineers

80 Rational eficiency of a powerplant

Hence, the maximum net work rejected is

(647.9 - 300) - 300 In 300

This is a significant increase over the rejected potential work for the ideal Joule cycle. However, in addition to the inefficiency of the turbine increasing the maximum net work rejected, it also increases the unavailable energy by the irreversibility, To(~4 - s ~ ~ ) , as shown in Fig 4.6(b). This is equal to

To(~4 - ~ 4 ~ ) = 300 x (0.7734 - 0.5328) = 72.2 kJ/kG

Since the maximum net work is the same for this case as the ideal one then the net work is given by

wnet = Gmt - [Gnct]E,jccd - To(~4 - ~ 4 ~ ) = 336.3 - 117.4 - 72.2 = 146.7 kJ/kg

This is the same value as found from the more basic calculation. The rational efficiency of this cycle is

146.7 W",t

&,,et 336.3 - 0.436 B R - --=--

(b) Compressor isentropic efficiency, qc = 80%. The work done in the compressor is

- 509.9 kJ/ kg w c = - - WCi,"

BC

Temperature after compressor, at 2, is

The energy added to the cycle is

q23 = cP(T3 - Tz) = 1 .0045 x (1200 - 807.6) = 394.2 kJ/kg

The net work from the cycle is wmt = wT + wc = 693.2 + (-509.9) = 183.3 kJ/kg, and the thermal efficiency is

wnct 183.3 - 0.465. V t h = - = - -

q23 394.2

The maximum net work output is defined by

GXt= b, - b,

Since both 2 and 3 are at the same pressure

b3 - b, = cp[(T3 - T,) - Toln -

Page 97: Advanced Thermodynamics for Engineers

Examples 81

Thus

(1200 - 807.6) - 3001n - = 273.6kJ/kg 807.6 1200 1

The rational efficiency of the cycle based on these dead state conditions is

w,, 183.3 it,,, 273.6

VR=-=-- - 0.670

The net work output of the device is made up of

wnet = $Et - - TO(s2 - sZs) = 273.6 - 50.9 - 40.2 = 182.5 W/kg

In this case, the inefficiency of the compressor has introduced a quantity of unavailable energy T0(s2 - s2J, which is depicted in Fig 4.7.

2 /// \ \ \ \ \ ,'

I ' 4

4s

Entropy, S

Fig. 4.7 The effect of inefficiency in the compressor

(c) Compressor isentropic efficiency vc = 80%, and turbine isentropic efficiency l;lT = 80%. This cycle combines the two inefficiencies considered above. Hence, the net work

output of the cycle is

w,, = wT + W C = 554.6 + (-509.9) = 44.7 kJ/kg

Net heat addition q23 = 394.2 kJ/kg. Hence, thermal efficiency is

wnet 44.7 - 0.113 47th = - = - -

q23 394.2

The net work output of this cycle is made up in the following way:

Wnet = $net - [$netlrejecred - - s4s) - - s 2 s )

= 273.6 - 117.4 - 72.2 - 40.2 = 43.8 kJ/kg

This equation shows how the inefficiencies reduce the net work obtainable from the cycle. However, an additional quantity of net work has been lost which is not evident

Page 98: Advanced Thermodynamics for Engineers

82

from the equation, and that is the reduction in maximum net work caused by the inefficiency of the compressor. The latter causes the temperature after compression to be higher than the isentropic value, and hence less fuel is necessary to reach the maximum cycle temperature.

4.5 Concluding remarks

The concept of a Second Law, or rational, efficiency has been introduced. This provides a better measure of how closely a particular power-producing plant approaches its maximum achievable efficiency than does the conventional ‘First Law’ thermal efficiency.

It has been shown that devices in which the working fluid changes phase (e.g. steam plant) can achieve rational efficiencies of 100 per cent, whereas those which rely on a single phase fluid (e.g. gas turbines) can never approach such a high value. The effect of irreversibility on rational efficiency has also been shown.

PROBLEMS

1 In a test of a steam powerplant (Fig P4.1), the measured rate of steam supply was 7.1 kg/s when the net rate of work output was 5000 kW. The condensate left the condenser as saturated liquid at 38°C and the superheated steam leaving the boiler was at 10 bar and 300°C. Neglecting the change in state of the feed water in passing through the feed pump, and taking the environment temperature as being equal to the saturation temperature of the condensate in the condenser, calculate the rational efficiency of the work-producing steam circuit (system A). Also calculate the thermal efficiency of the cyclic plant.

Rational efficiency of a powerpiant

I

Fig. P4.1 Steam turbine Dowemlant - [83.3%; 24.3%]

2 Again neglecting the change in state of the feed water in passing through the feed pump, calculate the thermal efficiency of an ideal Rankine cycle in which the

Page 99: Advanced Thermodynamics for Engineers

Problems 83

conditions of the fluid at inlet to and exit from the boiler are the same as those of the plant in Q.l. Hence confirm that the efficiency ratio of that plant is equal to the rational efficiency of the work-producing steam circuit calculated in Q. 1.

[29.2%; 83.3%]

3 Neglecting the temperature rise of the feed water in passing through the feed pump, calculate the mean temperature of heat reception in the ideal Rankine cycle of Q.1. Hence make a second, alternative calculation of the thermal eficiency of this ideal Rankine cycle.

[440 K; 29.2%]

4 Air enters the compressor of a simple gas turbine at a pressure of 1 bar and a temperature of 25°C. The compressor has a pressure ratio of 15 and an isentropic efficiency of 85%, and delivers air to a combustion chamber which is supplied with methane ( C q ) at 25°C. The products of combustion leave the chamber at 1450 K and suffer a pressure loss of 5% in passing through it. They are then expanded in a turbine with an isentropic efficiency of 90% to a pressure of 1.05 bar. Calculate (a) the air-fuel ratio of the engine; (b) the power output per unit mass flow of air, and (c) the rational efficiency of the engine. What is the maximum work per unit mass of exhaust gas that could be obtained if the ambient conditions are 1 bar and 25"C?

The enthalpy of reaction based on 25°C is -50 000 W/kg, and the Gibbs energy of reaction is -51 OOO kJ/kg. The products of combustion can be treated as a perfect gas with cp = 1.2 kJ/kg, and K = 1.35.

[0.02; 372.9 kJ/kg; 36.6%; 260.4 kJ/kg]

5 A steam turbine operates on a superheated Rankine cycle. The pressure and temperature of the steam leaving the boiler are 10 bar and 350°C respectively. The specific steam consumption of the plant is 4.55 kg/kWh. The pressure in the condenser is 0.05 bar.

If the feed pump work may be neglected, calculate the thermal efficiency of the plant, the turbine isentropic efficiency, and evaluate the rational efficiency. Also calculate the mean temperature of reception of heat in the boiler and use this in conjunction with the condenser temperature to evaluate the thermal efficiency. Explain why the value calculated by this method is higher than that obtained previously.

[26.5%; 86%; 84.84%; 30.85%]

6 Figure P4.6 depicts a closed cycle gas turbine operating on the Joule cycle (i.e. constant pressure heat addition and rejection, and isentropic compression and expansion). Energy is added to the working fluid (air) by a heat exchanger maintained at 1250 K, and rejected to another heat exchanger maintained at 300 K. The maximum temperature of the working fluid is 1150 K and its minimum temperature is 400 K. The pressure ratio of the compressor is 5 : 1.

Evaluate the irreversibilities introduced by the heat transfer processes and calculate

work ouput

energy addition (a) the First Law efficiency 91 =

Page 100: Advanced Thermodynamics for Engineers

84 Rational efficiency of a powerplant

work output availability of energy addition

(b) the Second Law efficiency 711 =

Assume cp = 1.005 H/kg K, 1c = 1.4 and the specific gas constant, R = 0.287 H/kg K.

Heat addition reservoir

1250 K b

Electric generator

Heat rejection reservoir

Fig. P4.6 Closed cycle gas turbine

Calculate the maximum efficiency that could be achieved from this system by

[36.9%; 48.5%] modification of the heat exchangers.

Page 101: Advanced Thermodynamics for Engineers

5 Efficiency of Heat Engines at Max imum

Power

5.1 maximum power output

The thermal efficiency of a Carnot cycle operating between high temperature ( T H ) and low temperature (T,) reservoirs is given by

Efficiency of an internally reversible heat engine when producing

This cycle is extremely idealised. It requires an ideal, reversible heat engine (internally reversible) but, in addition, the heat transfer from the reservoirs is also reversible (externally reversible). To achieve external reversibility it is necessary that the tempera- ture difference between the reservoirs and the engine is infinitesimal, which means that the heat exchanger surface area must be very large or the time to transfer heat must be long. The former is limited by size and cost factors whilst the latter will limit the actual power output achieved for the engine. It is possible to evaluate the maximum power oufpuf achievable from an internally reversible (endoreversible) heat engine receiving heat irreversibly from two reservoirs at TH and T,. This will now be done, based on Bejan (1988).

Assume that the engine is a steady-flow one (e.g. like a steam turbine or closed cycle gas turbine): a similar analysis is possible for an intermittent device (e.g. like a Stirling engine). A schematic of such an engine is shown in Fig 5.1.

The reservoir at TH transfers heat to the engine across a resistance and it is received by the engine at temperature T,. In a similar manner, the engine rejects energy at T2 but the cold reservoir is at T,. It can be assumed that the engine itself is reversible and acts as a Carnot cycle device with

This thermal efficiency is less than the maximum achievable value given by eqn (5.1) because T,/T, > T,/T,. The value can only approach that of eqn (5.1) if the temperature drops between the reservoirs and the engine approach zero.

Page 102: Advanced Thermodynamics for Engineers

86 Eficiency of heat engines at maximum power

Fig. 5.1 Internally reversible heat engine operating between reservoirs at TH and T,

The heat transfer from the hot reservoir can be defined as

Q H = uHAH(TH - T I )

where U H = heat transfer coefficient of hot reservoir (e.g. kW/m2 K)

and AH = area of heat transfer surface of hot reservoir (e.g. m2) QH = rate of heat transfer (e.g. kW).

(5.3)

The heat transfer to the cold reservoir is similarly

By the first law Qc = UcAc(T2 - Tc) (5.4)

w = Q H - QC (5.5) Now, the heat engine is internally reversible and hence the entropy entering and leaving it must be equal i.e.

This means that

It is possible to manipulate these equations to give W in terms of TH, Tc, UHAH, and the ratio T2/Tl = z. From eqn (5.4)

and, from eqn (5.6),

Page 103: Advanced Thermodynamics for Engineers

Eficiency of an internally reversible heat engine 87

Hence

Rearranging gives

and hence

w

t l+- ( 2:;) (5.10)

(5.11)

Thus the rate of work output is a function of the ratio of temperatures of the hot and cold reservoirs, the ratio of temperatures across the engine and the thermal resistances. The optimum temperature ratio across the engine (t) to give maximum power output is obtained when

- 0 a w at --

Differentiating eqn (5.1 1) with respect to t gives

a w 1 (1 - t) (t - TJT,) - (t - TJT,) f1 - t) 7. t2

1+-

Hence aw/& = o when t = 00 or t2 = Tc/TH. Considering only the non-trivial case, for maximum work output

(5.12)

This result has the effect of maximising the energy flow through the engine while maintaining the thermal efficiency (qh = 1 - T2/T1) at a reasonable level. It compromises between the high efficiency ( qh = 1 - Tc/TH) of the Carnot cycle (which produces zero energy flow rate) and the zero efficiency engine in which T2 = TI (which produces high energy flow rates but no power).

Page 104: Advanced Thermodynamics for Engineers

88 EfSiciency of heat engines at maximum power

Hence the efficiency of an internally reversible, ideal heat engine operating at maximum power output is

An example will be used to show the significance of this result.

Example Consider a heat engine is connected to a hot reservoir at 1600 K and a cold one at 400 K. The heat transfer conductances (UA) are the same on both the hot and cold sides. Evaluate the high and low temperatures of the working fluid of the internally reversible heat engine for maximum power output; also calculate the maximum power.

Solution Equation (5.11) gives the work rate (power) as

For maximum power output z = (TC/TH)'/' = (400/1600)1/2. Then

(1 - -) = - - - 200 units -- w (i-t) 1 1600

UHAH - 1600

1 2 8 - (1 + 1) 2

w and, from eqn (5.10)

= 400 units -- - QH

UnAH U H A H ( ~ -

Then

(Q, - W > = 200 units Qc 1 -- --=- Qc

UHAH UcAc &AH Thus

T , = ~ [ ~ + T c } = 2 x ( 2 0 0 + 4 0 0 ) = z UcAc 1200K

and 1200

2 T2 = - = 600 K

This results in the temperature values shown in Fig 5.2 for the engine and reservoirs.

Page 105: Advanced Thermodynamics for Engineers

Efficiency of an internally reversible heat engine 89

7' QH = 400 units

1' Q, = 200 units

Fig. 5.2 Example of internally reversible heat engine operating between reservoirs at TH = 1600 K and T, = 400 K with U,A,/U,A, = 1

The efficiency of the Camot cycle operating between the reservoirs would have been 7;lm = 1 - 400/1600 = 0.75 but the efficiency of this engine is 7;lm = 1 - 600/1200 = 0.50. Thus an engine which delivers maximum power is si&icantly less efficient than the Camot engine.

It is possible to derive relationships for the intermediate temperatures. Equation (5.9) gives

and eqn. (5.8) defines T2 as

T2 = - Qc +Tc UC Ac

Also Q, = Q, - W , and then

w Qc - - Q, -

Qc - (t - TCIT,)

UnAn TH UHAH TH UUAH TH Substituting from eqns (5.10) and (5.11) gives

- UH AH TH

(l+Z)

( 22) Substituting in eqn (5.8) gives

UHA, T , (Z - TCIT,) T2 = + Tc

UcAc l+-

U,A, T, (Z - t2)

= Tii2{ U, AH T;'~ + U,

+ Tc - - UcAc + &AH

I UHAH + UCAC

(5.8)

(5.13)

(5.14)

Page 106: Advanced Thermodynamics for Engineers

90 EfJiciency of heat engines at maximum power

Similarly

T u A T ’ / ~ + U~A,T$~

Un An -+ UcAc (5.15)

To be able to compare the effect of varying the resistances it is necessary to maintain the total resistance to heat transfer at the same value. For example, let

UHAH + UcAc = 2

Then, if UHAH/UcAc = 1 (as in the previous example), UHAH = 1.

UnAJUcAc = 112. This gives Consider the effect of having a high resistance to the high temperature reservoir, e.g.

Then

Then

Q c - -- UcAc

1600 (’”-- 1/4) (1 - 1/2) = 267 units, giving W = 177.8 units (1 + 1/2)/2

= 534 units and Q, = 355.6 units w

uti AH (1 -

(QH - W ) = 133.5 units, giving Qc = 177.8 units Qc *L- 1

- - UHAH 2 ~ U H A H

Hence

1 T, = - (133.5 + 400) = 2 x 533.5 = 1067K, and T , = 533.5 K.

z

If the resistance to the low temperature reservoir is high, i.e. U,AH/UcAc = 2, the situation changes, as shown below. First,

2 4

1+1/2 3/2 3 - -

2 UnAn = ~ - - - -

giving

= 133 units (1/2 - 1/4) (1 - 1/2) -- w - 1600

UnAn (1 + 2)/2

which results in W=177.8 kW and

-- Qn - 267 units, or Q = 355.6 kW UHAH

Page 107: Advanced Thermodynamics for Engineers

EfSiciency of an internally reversible heat engine 91

Then 2

-- ( QH - W ) = 267 units, giving Qc = 200 units Qc 2Qc -- UcAc UHAH UHAH

Hence

T I = ' [ A + T c ] = 2 ( 2 6 7 + 4 W ) = 1334K,andT2=667K UcAc

These results are shown graphically in Fig 5.3.

T, 120c

T* 600 T

1067 & K

533 k K

133

66;

T, = 400 K T Fig. 5.3 The effect of heat transfer parameters on engine temperatures for a heat engine operating

between reservoirs at TH = 1600 K and Tc = 400 K

Comparing the power outputs based on the same total resistance, i.e. UHAH + UcAc = 2, gives the following table.

UHAHluCAc Q H W Qc

1 400 200 200 112 355.6 177.7 177.7 2 355.6 177.7 177.7

It can be seen that the optimum system is one in which the high and low temperature resistances are equal. In this case the entropy generation (per unit of work) of the universe is minimised, as shown in the table below.

UHAHIUCAC QJTH QHIT, ASHIW QcIT, QcITc AScIW ZASlW

1 -0.25 +0.333 +4.17 x -0.333 + O S +4.17 x +8.34 x 112 -0.225 +0.333 +6.23 x -0.333 0.4443 +6.24 x +0.00125 2 -0.225 +0.266 +2.48 x -0.266 0.4443 +10.00 x +0.00125

Page 108: Advanced Thermodynamics for Engineers

92 EfSiciency of heat engines at maximum power

5.2 Efficiency of combined cycle internally reversible heat engines when producing maximum power output One way of improving the overall efficiency of power production between two reservoirs is to use two engines. For example, a gas turbine and steam turbine can be used in series to make the most effective use of the available temperature drop. Such a power plant is referred to as a combined cycle gas turbine, and this type of generating system was introduced in Chapter 3 in connection with pinch technology. These plants can be examined in the following way, based on the two heat engines in series shown in Fig 5.4.

Fig. 5.4 Two internally reversible engines in series forming a combined cycle device operating between two reservoirs at TH and T,

In this case the product UA will be replaced by a ‘conductivity’ C to simplify the notation. Then

Also, for E,

(5.16) (5.17) (5.18)

(5.19)

and for E,

Let z,

(5.20) _ - Q 2 Qc -- 7 . 3 7-4

= T2/T, and t2 = T4/T,; then rearranging eqns (5.16), (5.17) and (5.18) gives

T2 = Q2/c2 i- T3

= TH - QH/CH (5.21) (5.22)

Page 109: Advanced Thermodynamics for Engineers

Eficiency of combined cycle internally reversible heat engine 93

and

T4 = Q C K C + Tc

Also

and

Similarly

From eqns (5.16) and (5.26)

1 Q 1 Q c Q H = C n T H - - L + - -+Tc [ T I ( C2 t 2 { T c })] Rearranging eqn (5.27) gives

which can be written as

The power output of engine E , can be obtained from eqn (5.29) as

In a similar manner Wc/CHTH can also be evaluated as

(5.23)

(5.24)

(5.25)

(5.26)

(5.27)

(5.28)

(5.29)

(5.30)

(5.31)

(5.32)

Page 110: Advanced Thermodynamics for Engineers

94 EfSiciency of heat engines at maximum power

It can be seen from eqns (5.30) and (5.32) that it is possible to split the power output of the two engines in an arbitrary manner, dependent on the temperature drops across each engine. The ratio of work output of the two engines is

w, 1 (1 -rl)

w c r1 (1 -r2) - (5.33)

This shows that if the temperature ratios across the high temperature and the low temperature engines are equal (i.e. t, = r2 ) , the work output of the low temperature engine will be

ci/, = T1WH (5.34)

Hence the work output of the low temperature engine will be lower than that of the high temperature engine for the same temperature ratio. The reason comes directly from eqns (5.30) and (5.31), which show that the work output of an engine is directly proportional to the temperature of the ‘heat’ at entry.

The power output of a combined cycle power plant is the sum of the power of the individual engines, hence

This may be reduced to

wcc CH TH

- t,r21

The efficiency of the combined cycle is defined by

- 1 - vth=-- wcc

Q H

Equation (5.36) shows that the expression for

(5.35)

(5.36)

the efficiency of the combined cycle engine at maximum power output is similar to that for the efficiency of a single heat engine, except that in this case the temperature ratio of the single engine is replaced by the product of the two temperature ratios. If the combined cycle device is considered to be two endoreversible heat engines connected by a perfect conductor (i.e. the resistance between the engines is zero; C, = m) then T3 = T,, and eqn (5.36) becomes

(5.37)

The efficiency given in eqn (5.37) is the same efficiency as would be achieved by a single endoreversible heat engine operating between the same two temperature limits, and what would be expected if there was no resistance between the two engines in the combined cycle plant.

Page 111: Advanced Thermodynamics for Engineers

Eflciency of combined cycle internally reversible heat engine 95

To determine the efficiency of the combined cycle plant, composed of endoreversible heat engines, producing maximum power output requires the evaluation of the maxima of the surface WE plotted against the independent variables z1 and t2. It is difficult to obtain a mathematical expression for this and so the maximum work will be evaluated for some arbitrary conditions to demonstrate the necessary conditions. This will be done based on the following assumptions:

temperatures: TH = 1600 K, conductivities: CH = C2 = Cc = 1

T, = 400 K

1' Qn AI Qn

'' Q i $c

T3 -

T4 - r

l' Qc

I' Qc

Page 112: Advanced Thermodynamics for Engineers

96 Eficiency of heat engines at maximum power

While these assumptions are arbitrary, it can be shown that the results obtained are logical and general. The variation of maximum power output with temperature ratios across the high and low temperature engines is shown in Fig 5.5. It can be seen that the maximum power occurs along a ridge which goes across the base plane. Examination shows that, in this case, this obeys the equation zlzz = 0.5 = .ITc'/7;1. Hence, the efficiency of a combined cycle heat engine operating at maximum power output is the same as the efficiency obtainable from a single heat engine operating between the same two reservoirs. This solution is quite logical, otherwise it would be possible to arrange heat engines as in Fig 5.6, and produce net work output while transferring energy with a single reservoir.

Thus the variation of z1 with z2 to produce maximum power output is shown in Fig 5.7: all of these combinations result in the product being 0.5.

2 06

6 055 ' 0 5 ' 0 5 055 06 065 0 7 075 0 8 085 09 095 1

Temperature ratio of high temperature engine

Fig. 5.7 Variations in temperature ratio of high and low temperature heat engines, which produce maximum power output

5.3 Concluding remarks

A new method for assessing the potential thermal efficiency of heat engines has been introduced. This is based on the engine operating at its maximum power output, and it is shown that the thermal efficiency is significantly lower than that of an engine operating on a Carnot cycle between the same temperature limits. The loss of efficiency is due to external irreversibilities which are present in all devices producing power output.

It has been shown that a combined cycle power plant cannot operate at a higher thermal efficiency than a single cycle plant between the same temperature limits. However, the use of two cycles enables the temperature limits to be widened, and thus better efficiencies are achieved.

PROBLEMS

1 Explain why the Carnot cycle efficiency is unrealistically high for a real engine. Introducing the concept of external irreversibility, evaluate the efficiency of an endoreversible engine at maximum power output.

Page 113: Advanced Thermodynamics for Engineers

Problems 97

Consider a heat engine connected to a high temperature reservoir at TH = 1200 K and a low temperature one at T, = 300 K. If the heat transfer conductances from the reservoirs to the engine are in the ratio (UA),/(UA), = CH/CC = 2, evaluate the following:

0 the maximum Carnot cycle efficiency; 0 the work output of the Carnot cycle; 0 the engine efficiency at maximum power output; 0 the maximum power output, W / C ~ ; 0 the maximum and minimum temperatures of the working fluid at maximum power

Derive all the necessary equations, but assume the Carnot efficiency, 17 = 1 - T,/TH.

output.

[0.75; 0; 0.5; 100; 1000 K; 500 K]

A heat engine operates between two finite reservoirs, initially at 800 K and 200 K respectively. The temperature of the hot reservoir falls by 1 K for each kJ extracted from it, while the temperature of the cold reservoir rises by 1 K for each kJ added. What is the maximum work output from the engine as the reservoir temperatures equalise? Is the equalisation temperature for maximum work a higher or lower limit of the equalisation temperature?

[200 kJ; lower]

Closed cycle gas turbines operate on the infernally reversible Joule cycle with an efficiency of

where rp = pressure ratio of the turbine, and K = ratio of specific heats, c,/c, = 1.4. This equation significantly overestimates the efficiency of the cycle when external

irreversibilities are taken into account. Shown in Fig P5.3 is a T-s diagram for a closed

I, ,,’p2 = 12 bar

_,*’ p, = 1 bar

Entropy, S

Fig. P5.3 Temperature-entropy diagram for Joule cycle

Page 114: Advanced Thermodynamics for Engineers

98 Eficiency of heat engines at maximum power

cycle gas turbine receiving energy from a high temperature reservoir at TH = 1200 K and rejecting energy to a low temperature reservoir at T, = 400 K.

(i) Evaluate the Carnot efficiency and compare this with the Joule efficiency: explain why the Joule efficiency is the lower.

(ii) Calculate the ratio of the gas turbine cycle work to the energy delivered from the high temperature reservoir (QH for the Carnot cycle). This ratio is less than the Joule efficiency - explain why in terms of unavailable energy.

(iii) Calculate the external irreversibilities and describe how these may be reduced to enable the Joule efficiency to be achieved.

(iv) What are the mean temperatures of energy addition and rejection in the Joule cycle? What would be the thermal efficiency of a Carnot cycle based on these mean temperatures?

[0.667; 0.508; 0.4212; ZH/cP=79.9; 994 K; 489 K]

4 Explain why the Carnot cycle overestimates the thermal efficiency achievable from an engine producing power output. Discuss why external irreversibility reduces the effective temperature ratio of an endoreversible engine.

Show that the thermal efficiency at maximum power output of an endoreversible engine executing an Otto cycle is

where TH = maximum temperature of the cycle and T, = minimum temperature of the cycle.

5 The operating processes of a spark-ignition engine can be represented by the Otto cycle, which is internally reversible and gives a thermal efficiency of

where r = volumetric compression ratio; K = ratio of specific heats, cp/c,. An Otto cycle is depicted in Fig P5.5, and the temperatures of the two reservoirs

associated with the cycle are shown as TH and T,. The thermal efficiency of a Carnot cycle operatirig between these twc. reservoirs is q = 1 - T,/TH. This value is signifi- cantly higher than that of the Otto cycle operating between the same reservoirs. Show the ratio of net work output for the Otto cycle to the energy transferred from the high temperature reservoir for the Carnot cycle, QH, is

where T2 and T3 are defined in Fig P5.5.

Page 115: Advanced Thermodynamics for Engineers

Problems 99

Entropy, S

Fig. P5.5 Temperature-entropy diagram for Otto cycle

Explain why this value of differs from that of the Otto cycle, and discuss the significance of the term

T3 - T2 T3 In ( TdT2 )

Evaluate the entropy change required at the high temperature reservoir to supply the Otto cycle in terms of the entropy span, s3 - s2, of the Otto cycle.

[As, = C , T ~ ( ~ * ~ ~ ~ ~ - l)/&i]

6 It is required to specify an ideal closed cycle gas turbine to produce electricity for a process plant. The first specification requires that the turbine produces the maximum work output possible between the peak temperature of 1200 K and the inlet tempera- ture 300 K. The customer then feels that the efficiency of the turbine could be improved by

(i) incorporating a heat exchanger; (ii) introducing reheat to the turbine by splitting the turbine pressure ratio such

that the pressure ratio of each stage is the square root of the compressor pressure ratio.

Evaluate the basic cycle efficiency, and then evaluate separately the effects of the heat exchange and reheat. If these approaches have not increased the efficiency, propose another method by which the gas turbine performance might be improved (between the same temperature limits), without reducing the work output; evaluate the thermal efficiency of the proposed plant. The ratio of specific heats may be taken as K = 1.4.

[OS; 0.4233; 0.5731 ]

Page 116: Advanced Thermodynamics for Engineers

General Thermodynamic Relationships: single component systems, or systems of

constant com position

The relationships which follow are based on single component systems, or systems of constant composition. These are a subset of the more general equations which can be derived, and which allow for changes in composition. It will be shown in Chapter 12 that if a change in composition occurs then another term defining the effect of this change is required.

6.1 The Maxwell relationships

The concept of functional relationships between properties was introduced previously. For example, the Second Law states that, for a reversible process, T , s, u, p and v are related in the following manner

T d S = d U + p dV (6.1) or, in specific (or molar) terms

T d s = d u + p d v (6.la)

Rearranging eqn (6.la) enables the change of internal energy, du, to be written

d u = T ds-p dv (6.2)

(6.2a)

It will be shown in Chapter 12 that, in the general case where the composition can change, eqn (6.1) should be written

TdS = dU + p dV - C p i d n , 1

where

p , = chemical potential of component i n, = amount of component i

(6.lb)

The chemical potential terms will be omitted in the following analysis, although similar equations to those below can be derived by taking them into account.

Page 117: Advanced Thermodynamics for Engineers

The Maxwell relationships 101

It can be seen from eqns (6.2) and (6.2a) that the specific internal energy can be represented by a three-dimensional surface based on the independent variables of entropy and specific volume. If this surface is continuous then the following relationships can be based on the mathematical properties of the surface. The restriction of a continuous surface means that it is ‘smooth’. It can be seen from Fig 6.1 that the p-v- T surface for water is continuous over most of the surface, but there are discontinuities at the saturated liquid and saturated vapour lines. Hence, the following relationships apply over the major regions of the surface, but not across the boundaries. For a continuous surface z = z ( x , y) where z is a continuous function. Then

d z = - dx+ - dy (6.3) (a (3 Let

I V = ( $ ) ~ and .=(:)I (6.4)

Then

dz = M dx + N dy (6.5) For continuous functions, the derivatives

aZz d2Z

ax ay ay ax and -

Page 118: Advanced Thermodynamics for Engineers

102 General thermodynamic relationships

are equal, and hence

(?)x=(:)y

Consider also the expressions obtained when z = z (x , y ) and x and y are themselves related to additional variables u and v, such that x = x(u , v) and y = y ( u , v). Then

Let z = v, and u = x, then x = x ( z ) and

(2) V =0, and ($)v=l

Hence

giving

(6.7)

These expressions will now be used to consider relationships derived previously. The

d u = T d s - p d v (6.10) dh = T ds + v dp (6.1 1) df = -p dv - s dT (6.12) dg = v dp - s dT (6.13)

following functional relationships have already been obtained

Consider the expression for du, given in eqn (6.10) then, by analogy with eqn (6.3)

T = ( $)v, -p = (%), and (5). = -( $)" (6.14)

In a similar manner the following relationships can be obtained, f o r constant composition or single component systems

T=($); ($)s=($) P

-P=(:); -s=($; ($)v=($)T

v=($); -.=($); ($)p=($)T

(6.15)

(6.16)

(6.17)

Page 119: Advanced Thermodynamics for Engineers

The Maxwell relationships 103

If the pairs of relationships for T are equated then

(z)v=($)p and, shilarly

In addition to these equivalences, eqns (6.14) to (6.17) also show that

($l=

($)p = -MT Equations (6.19) are called the Maxwell relationships.

(6.18a)

(6.18b)

(6.18~)

(6.18d)

(6.19a)

(6.19b)

(6.19~)

(6.19d)

6.1.1 GRAPHICAL-. INTERPRETATION OF MAXWELL RELATIONS

Let a system comprising a pure substance execute a small reversible cycle 1-2-3-4-1 consisting of two isochors separated by dV and two isentropes separated by ds; these cycles are shown in Fig 6.2.

If the cycles are reversible, the area on the T-s diagram must equal the area on the p - v diagram. Now the difference in pressure between lines 1 and 2 is given by

and to the first order the difference in pressure between lines 3 and 4 is the same. Hence

the area on the p - v diagram = - ( 3 ds dv (6.20a)

Page 120: Advanced Thermodynamics for Engineers

104 General thermodynamic relationships

e,

2 m

Y n 5

2-,' ,

;;/''' v + 6v *-.-*\--... 4 - - . ... . / + 4

* * > 2 E

Applying the same approach to the T-s diagram, the difference in temperature between lines 1 and 4 is

-($) dv S

(6.20b)

which is negative because the temperature decreases as the volume increases, and the area of the diagram is

-( $) dv ds S

Thus, equating the two areas gives

-( n), dv ds = ($) ds dv V

(6.21)

and hence

6.2

When performing certain types of calculation it is useful to have data on the values of the specific heat capacities of the substance under consideration and also the variation of these specific heat capacities with the other properties. The specific heat capacities themselves are first derivatives of the internal energy ( u ) and the enthalpy ( h ) , and hence the variations of specific heat capacities are second derivatives of the basic properties.

Uses of the thermodynamic relationships

By definition, the specific heat capacity at constant volume is

(6.22)

Page 121: Advanced Thermodynamics for Engineers

But

($), = T, and thus c, = T - (3 Similarly

Uses of the thermodynamic relationships 105

(6.23)

(6.24)

Consider the variation of the specific heat capacity at constant volume, c,, with specific volume, v, if the temperature is maintained constant. This can be derived in the following way:

Now, from eqn (6.19~)

giving

(6.25)

Hence, if data are available for a substance in the form of a p-v-T surface, or a mathematical relationship, it is possible to evaluate (a2p/aT2), and then (dc,/av),. By integration it is then possible to obtain the values of c, at different volumes. For example, consider whether the specific heat capacity at constant volume is a function of the volume for both an ideal gas and a van der Waals’ gas.

Ideal gas The gas relationship for an ideal gas is

pv = RT

and

($)”=: and ($)“=O

(6.26)

(6.27)

Hence c ,+ f (v ) for an ideal gas. This is in agreement with Joule’s experiment for assessing the change of internal energy, u, with volume.

van der Waals’ gas The equation of state of a van der Waals’ gas is

RT a p = - - -

v - b v2 (6.28)

Page 122: Advanced Thermodynamics for Engineers

106 General thermodynamic relationships

where a and b are constants. Hence

R and ($)=O (6.29)

Again c, # f( v ) for a van der Waals’ gas. So, for ‘gases’ obeying these state equations c , + f ( v ) , but in certain cases and under

certain conditions c, could be a function of the volume of the system and it would be calculated in this way. Of course, if these equations are to be integrated to evaluate the internal energy it is necessary to know the value of c, at a datum volume and temperature.

Similarly, the variation of the specific heat capacity at constant pressure with pressure can be investigated by differentiating the specific heat capacity with respect to pressure, giving

(6.30)

This equation can be used to see if the specific heat capacity at constant pressure of gases obeying the ideal gas law, and those obeying van der Waals’ equation are functions of pressure. This is done below.

Ideal gas

p v = RT

Hence

(6.31)

This means that the specific heat capacity at constant pressure for a gas obeying the ideal gas law is not a function of pressure, i.e. cp # f( p ) for an ideal gas. This conclusion is in agreement with the Joule-Thomson experiment for superheated gases.

Van der Waals’ gas

RT a p = - - -

v - b v 2

Equation (6.28) can be rewritten as

( p + ; ) ( v - b ) = R T (6.32)

which expands to

a ab

V V 2 p v + - - p b - - = RT

Page 123: Advanced Thermodynamics for Engineers

Uses of the thermodynamic relationships 107

Differentiating implicitly gives

which can be rearranged to give

( $ ) J P - 7 + 7 } - , a 2ab

and hence

R

p-- a ( 'e) ($),= V 2

This can be differentiated again to give

which results in

a 2ab

giving

2a 6ab

, which, on substituting for a 2ab

($+=- 2a 6ab

(6.33)

(6.34)

Hence, for a van der Waals' gas, cp = f ( p ) . This means that allowance would have to be taken of this when evaluating the change of

specific heat capacity at constant pressure for a process in which the pressure changed. This can be evaluated from

cp = 5,4' - T( -), aZv dP aT2

(6.35)

where (a2v/aT2), can be calculated from eqn (6.34).

Page 124: Advanced Thermodynamics for Engineers

108 General thermodynamic relationships

6.3 Tds relationships

Two approaches will have been used previously in performing cycle calculations. When evaluating the performance of steam plant and refrigeration equipment, much emphasis was placed on the use of tables, and if the work done between two states was required, enthalpy values at these states were evaluated and suitable subtractions were performed. When doing cycle calculations for gas turbines and internal combustion engines the air was assumed to be an ideal gas and the specific heat capacity was used. The most accurate way of performing such calculations is, in fact, to use the enthalpy or internal energy values because these implicitly integrate the specific heat capacity over the range involved. The following analyses will demonstrate how these values of internal energy and enthalpy are obtained.

The two-property rule can be used to define entropy as the following functional relations hip

s = S(V, T) (6.36)

Hence

Now, by definition

and from the Maxwell relationships, eqn (6.19c),

Hence

(6.37)

(6.38)

This is called the first Tds relationship.

entropy is a function of temperature and pressure, then The second Tds relationship can be derived in the following way. If it is assumed that

s = s (T,p) (6.39)

and

(6.40)

By definition,

cp = T( $) P

Page 125: Advanced Thermodynamics for Engineers

Tds relationships 109

and, from eqn (6.19d)

(:)T = -( &), Thus, the second Tds relationship is

The third Tds relationship is derived by assuming that

s = s ( p , v)

Now, by definition,

A similar expression can be derived for c,, and this enables the equation

(6.41)

(6.42)

(6.43)

to be obtained. This is the third Tds relationship. It is now possible to investigate the variation of internal energy and enthalpy with

independent properties for gases obeying various gas laws, The internal energy of a substance can be expressed as

(6.45) u = u (T, V )

Hence, the change in internal energy is

(6.46)

Now

cv = ($) and from the second law

d u = T ds-p dv (6.47)

Substituting for Tds in eqn (6.47) from the first Tds relation (eqn (6.38)) gives

(6.48)

Page 126: Advanced Thermodynamics for Engineers

1 10 General thermodynamic relationships

and hence eqn (6.46) becomes

or

(6.49)

(6.50)

Equation (6.50) can be used to evaluate the variation in internal energy with volume for both ideal and van der Waals’ gases.

Ideal gas

The equation of state of an ideal gas is pv = RT and hence

Thus

(6.51)

(6.52)

Hence, the specific internal energy of an ideal gas is not a function of its specific volume (or density). This is in agreement with Joule’s experiment that u +f(v) at constant temperature.

Van der Waals’ gas

The equation of state of a van der Waals’ gas is

RT a p=---

v - b v2

and hence

R

V - b

which gives the change of internal energy with volume as

( g ) T = [ 2 - P ] = ;

(6.53)

(6.54)

This means that the internal energy of a van der Waals’ gas is a function of its specific volume, or density. This is not surprising because density is a measure of the closeness of the molecules of the substance, and the internal energy variation is related to the force of attraction between the molecules. This means that some of the internal energy in a van der

Page 127: Advanced Thermodynamics for Engineers

Relationships between spec@ heat capacities 1 11

Waals' gas is stored in the attraction forces between the molecules, and not all of the thermal energy is due to molecular motion, as was the case for the ideal gas.

Now, if it is required to calculate the total change in internal energy, u, for a change in volume and temperature it is necessary to use the following:

For an ideal gas

(6.55)

i.e. for an ideal gas u = f(T).

done in the following way: It is also possible to use the following approach to show whether u = f( p ) . This may be

Now

(fl=. ($), = -;, P and sincep # 0,

Hence, for an ideal gas u # f( p ) . For a van der Waals' gas

a du = C , d T + - dv

V 2

Hence for a van der Waals' gas u = f(T, v).

(6.56)

6.4 Relationships between specific heat capacities

These relationships enable the value of one specific heat capacity to be calculated if the other is known. This is useful because it is much easier to measure the specific heat capacity at constant pressure, cp , than that at constant volume, c,. Using the two-property rule it is possible to write

s = s(T, V) (6.57)

and, if s is a continuous function of temperature and volume,

ds = ( $), dT + ( dv (6.58)

If eqn (6.58) is differentiated with respect to temperature, T, with pressure maintained constant, then

(6.59)

Page 128: Advanced Thermodynamics for Engineers

1 12 General thermodynamic relationships

The definitions of the specific heat capacities are

c, = T( $) and cp=T($-) V P

Hence

5c-5 =(?#I) P T

From Maxwell relations (eqn (6.19c)),

Thus

c P - c v -(*)(?) T aT , aT

The mathematical relationship in eqn (6.9)

can be rearranged to give

(%), =

Thus

(6.60)

(6.61)

(6.62)

(6.63)

Examination of eqn (6.64) can be used to define which specific heat capacity is the larger. First, it should be noted that T and (av/aT)i are both positive, and hence the sign of cP - C , is controlled by the sign of (ap/av),. Now, for all known substances (ap/av), is negative. If it were not negative then the substance would be completely unstable because a positive value means that as the pressure is increased the volume increases, and vice versa. Hence if the pressure on such a substance were decreased its volume would decrease until it ceased to exist.

Thus cp - c, must always be positive or zero, i.e. cpa c,. The circumstances when cp = c, are when T = 0 or when (ap/aT), = 0, e.g. at 4°C for water. It can also be shown that cP> c, by considering the terms in eqn (6.61) in relation to the state diagrams for substances, as shown in Fig 6.3. The term (ap/aT),, which can be evaluated along a line at constant volume, v, can be seen to be positive, because as the temperature increases, the pressure increases over the whole of the section. Similarly, (av/aT),, which can be evaluated along a line at constant pressure, p , is also positive. If both these terms are positive then cp z c,.

Page 129: Advanced Thermodynamics for Engineers

Relationships between specific heat capacities 1 13

\

Specific volume

Fig. 6.3 p - v section of state diagram showing isotherms

Now consider cp - c , for an ideal gas, which is depicted by the superheated region in Fig 6.3. The state equation for an ideal gas is

p v = RT

Differentiating gives

(?E)T= -- P V

and

(+Jp= 3 Thus

(6.65)

(6.66)

(6.67)

The definition of an ideal gas is one which obeys the ideal gas equation (eqn (6.26)), and in which the specific heat capacity at constant volume (or pressure) is a function of temperature alone, i.e. c, = f ( T ) . Hence, from eqn (6.67), cp = c, + R = f ( T ) + R = f ' ( T ) if R is a function of T alone. Hence, the difference of specific heat capacities for an ideal gas is the gas constant R . Also

1 and ($) = B v

P

Page 130: Advanced Thermodynamics for Engineers

1 14 General thermodynamic relationships

and thus

(6.68)

where = coefficient of expansion (isobaric expansivity)

k = isothermal compressibility. These parameters are defined in Chapter 7.

Expressions for the difference between the specific heat capacities, cp - e,, have been derived above. It is also interesting to examine the ratio of specific heat capacities, K = cp/c,. The definitions of cp and c, are

cv = T($), and cP=T($) ,

and thus the ratio of specific heat capacities is

c p = (as/aT),

as aT ap as aT av Now, from the mathematical relationship (eqn (6.9)) for the differentials,

(z)p(G)i%)T = - = (aT).o)(x)T

giving

cp =(n),o) (?E)(”), CV as aT ap

From the Maxwell relationships

aT aV =

Now, from the two-property rule, T = T( p , v ) and hence

(?!!)T= -(?E) aT , (”) aV Thus

(6.69)

(6.70)

(6.71)

(6.72)

(6.73)

(6.74)

Page 131: Advanced Thermodynamics for Engineers

The Clausius-Clapeyron equation 115

The denominator of eqn (6.74), ~ ( J p / a v ) , , is the reciprocal of the isothermal compressibility, k. By analogy, the numerator can be written as l/k,, where k, = adiabatic, or isentropic, compressibility. Thus

(6.75)

If the slopes of isentropes in the p - v plane are compared with the slopes of isotherms (see Fig 6.4), it can be seen that cp/cv > 1 for a gas in the superheat region.

Volume

Fig. 6.4 Isentropic and isothermal lines for a perfect gas

6.5 The Clausius-Clapeyron equation From the Maxwell relationships (eqn 6.19c),

(E), = ($jV The left-hand side of this equation is the rate of change of entropy with volume at constant temperature.

If the change of phase of water from, say, liquid to steam is considered, then this takes place at constant pressure and constant temperature, known as the saturation values. Hence, if the change of phase between points 1 and 2 on Fig 6.5 is considered,

(g)T= - which, on substituting from eqn (6.19c), becomes

(g)v= -

(6.76)

(6.77)

If there is a mixture of two phases, say steam and water, then the pressure is a function of temperature alone, as shown in Chapter 7. Thus

(6.78)

Page 132: Advanced Thermodynamics for Engineers

1 16 General thermodynamic relationships

Now (ah/as), = T, and during a change of phase both T and p are constant, giving h, - h, = T ( s , - SI). Thus

(6.79)

Specific volume

Fig. 6.5 Change of phase depicted on a p-v diagram

, Critical paint

Fig. 6.6 Evaporation processes shown on state diagrams: (a) p-v diagram; (b) T-s diagram

Page 133: Advanced Thermodynamics for Engineers

The Clausius -Clapeyron equation 117

This equation is known as the Clapeyron equation. Now h2 - h,, is the latent heat, hfg. If point 1 is on the saturated liquid line and point 2 is on the saturated vapour line then eqn (6.79) can be rewritten as

(6.80)

This can be depicted graphically, and is shown in Fig 6.6. If the processes shown in Figs 6.6(a) and (b) are reversible, then the areas are

equivalent to the work done. If the two diagrams depict the same processes on different state diagrams then the areas of the 'cycles' must be equal. Hence

(6.81) ( ~ 2 - V I ) dp = ( ~ 2 - SI) dT Now, the change of entropy is sfg = s2 - s, = hf,/T, again giving (eqn (6.80))

Equation (6.80) is known as the Clausius-Clapeyron equation.

6.5.1 THE USE OF THE CLAUSIUS-CLAPEYRON EQUATION

If an empirical expression is known for the saturation pressure and temperature, then it is possible to calculate the change of entropy and specific volume due to the change in phase. For example, the change of enthalpy and entropy during the evaporation of water at 120°C can be evaluated from the slope of the saturation line defined as a function of pressure and temperature. The values of entropy and enthalpy for dry saturated steam can also be evaluated if their values in the liquid state are known. An extract of the properties of water is given in Table 6.1, and values of pressure and temperature on the saturation line have been taken at adjacent temperatures.

Table 6.1 Excerpt from steam tables, showing values on the saturated liquid and vapour lines

Specific Change of Entropy of Enthalpy of liquid, Temperature Pressure volume specific liquid, sf h f /("C> /(bar) /(m3/kg) volume, Vfg /(kJ/kg K) /(kJ/kg K)

115 1.691 120 1.985 0.001060 0.8845 1.530 505 125 2.321

At 120°C the slope of the saturation line is

dp 2.321 - 1.691 _ - - dT 10

x lo5 = 6.30 x lo3 N/m2 K

From the Clausius-Clapeyron equation, eqn (6.80),

s ~ - s , dp sg-sf - -

vfg dT v f g

Page 134: Advanced Thermodynamics for Engineers

1 18 General thermodynamic relationships

giving

dP S, = sf + vfg - = 1.530 + 0.8845 x 6.30 x 103/103

dT = 1.530 + 5.57235 = 7.102 kJ/kgK

The specific enthalpy on the saturated vapour line is

h, = h, + Ts, = 505 + (120 + 273) x 5.57235 = 2694.9 kJ/kg K

These values are in good agreement with the tables of properties; the value of s, = 7.127 kJ/kgK, while h, = 2707 kJ/kg at a temperature of 120.2”C.

6.6 Concluding remarks Thermodynamic relationships between properties have been developed which are independent of the particular fluid. These can be used to evaluate derived properties from primitive ones and to extend empirical data.

PROBLEMS 1 For a van der Waals’ gas, which obeys the state equation

RT a p=---

v - b v 2

show that the coefficient of thermal expansion, B, is given by

Rv2(v - b)

RTv3 - 2a(v - b)2 B =

and the isothermal compressibility, k, is given by

v2(v - b)* RTv3 - 2a(v - b)’

k =

Also evaluate

for the van der Waals’ gas. From this find cp - c, for an ideal gas, stating any assumptions made in arriving at the solution.

2 Prove the general thermodynamic relationship

and evaluate an expression for the variation of cp with pressure for a van der Waals’

Page 135: Advanced Thermodynamics for Engineers

Problems 119

gas which obeys a law RT a

p = - - - 2 v , -b V,

where the suffix 'm' indicates molar quantities and R is the universal gas constant. It is possible to represent the properties of water by such an equation, where

m3 R = 8.3143 kJ/kmolK (m3)' bar

a = 5.525 ; b = 0.03042 -*

OunoD2 kmol '

Evaluate the change in cp as the pressure of superheated water vapour is increased from 175 bar to 200 bar at a constant temperature of 425°C. Compare this with the value you would expect from the steam tables abstracted below:

425 0.013914 3014.9 0.011458 2952.9 450 0.015174 3109.7 0.012695 3060.1

What is the value of (acP/ap), of a gas which obeys the Clausius equation of state, p = RT/(v, - b)?

3 Show that if the ratio of the specific heats is 1.4 then

(%),=;(%I 4 Showthat

5 Show

(a) Tds=c ,dT-T - dp (3

Page 136: Advanced Thermodynamics for Engineers

120 General thermodynamic relationships

6 Show that, for a pure substance,

(s)p=($)v+(:)E)p and from this prove that

cp - cv = T( *) (") aT a T

The following table gives values of the specific volume of water, in m3/kg. A quantity of water is initially at 30"C, 20 bar and occupies a volume of 0.2 m3. It is heated at constant volume to 50°C and then cooled at constant pressure to 30°C. Calculate the net heat transfer to the water.

Specific volume/(m3/kg)

Pressure/(bar) 20 200 Temperature/("C ) 30 0.0010034 0.0009956 50 0.00101 12 0.0010034

[438 kT]

7 Assuming that entropy is a continuous function, s = s(T, v), derive the expression for entropy change

T

and similarly, for s = s(T, p) derive

dT T ( l;)p ds=c , - - - dp

Apply these relationships to a gas obeying van der Waals' equation

RT a p = - - - v - b v 2

and derive an equation for the change of entropy during a process in terms of the basic properties and gas parameters. Also derive an expression for the change of internal energy as such a gas undergoes a process.

Page 137: Advanced Thermodynamics for Engineers

7 Equations of State

The properties of fluids can be defined in two ways, either by the use of tabulated data (e.g. steam tables) or by state equations (e.g. perfect gas law). Both of these approaches have been developed by observation of the behaviour of fluids when they undergo simple processes. It has also been possible to model the behaviour of such fluids from ‘molecular’ models, e.g. the kinetic theory of gases. A number of models which describe the relationships between properties for single component fluids, or constant composition mixtures, will be developed here.

7.1 Ideal gas law

The ideal and perfect gas laws can be developed from a number of simple experiments, or a simple molecular model. First the experimental approach will be considered.

If a fixed mass of a single component fluid is contained in a closed system then two processes can be proposed:

(i) the volume of the gas can be changed by varying the pressure, while maintaining the temperature constant;

(ii) the volume of the system can be changed by varying the temperature, while maintaining the pressure constant.

The first process is an isothermal one, and is the experiment proposed by Boyle to define Boyle’s law (also known as Mariotte’s law in France). The second process is an isobaric one and is the one used to define Charles’ law (also known as Gay-Lussac’s law in France).

The process executed in (i) can be described mathematically as

v = v( P I T (7.1)

while the second one, process (ii), can be written

v = v(T) ,

Since these processes can be undergone independently then the relationship between the three properties is

(7.3)

Page 138: Advanced Thermodynamics for Engineers

122 Equations of state

Equation (7.3) is a functional form of the equation ofstare of a single component fluid. It can be seen to obey the two-property rule, which states that any property of a single component fluid or constant composition mixture can be defined as a function of two independent properties. The actual mathematical relationship has to be found from experiment (or a simulation of the molecular properties of the gas molecules), and this can be derived by knowing that, if the property, v, is a continuous function of the other two properties, p and T, as discussed in Chapter 6, then

dv = ( $)T dp + ( $)p dT (7.4)

Hence, if the partial derivatives (av/ap) , and (av/aT), can be evaluated then the gas law will be defined. It is possible to evaluate the first derivative by a Boyle’s law experiment, and the second one by a Charles’ law experiment. It is found from Boyle’s law that

pv =constant

giving

P =- - V

Similarly, it is found from Charles’ law that

V _ - - constant T

giving

V

($)p =r Substituting eqns (7.6) and (7.8) into eqn (7.4) gives

V V dv = - - dp +- - dT

P T

which may be integrated to give

= constant T

(7.5)

(7.7)

(7.8)

(7.9)

(7.10)

Equation (7.10) is known as the ideal gas law. This equation contains no information about the internal energy of the fluid, and does not define the specific heat capacities. If the specific heat capacities are not functions of temperature then the gas is said to obey the perfect gas law: if the specific heat capacities are functions of temperature (Le. the internal energy and enthalpy do not vary linearly with temperature) then the gas is called an ideal gas.

It is possible to define two coefficients from eqn (7.4), which are analogous to concepts used for describing the properties of materials. The first is called the isothermal compressibility, or isothermal bulk modulus, k. This is defined as the ‘volumetric strain’

Page 139: Advanced Thermodynamics for Engineers

Van der Waals’ equation of state 123

(7.1 1)

The negative sign is introduced because an increase in pressure produces a decrease in volume and hence (av/ap) < 0; it is more convenient to have a coefficient with a positive value and the negative sign achieves this. The isothermal compressibility of a fluid is analogous to the Young’s modulus of a solid.

The other coefficient that can be defined is the coeficient of expansion, B. This is defined as the ‘volumetric strain’ produced by a change in temperature, giving

(7.12)

The coefficient of expansion is analogous to the coefficient of thermal expansion of a solid material.

Hence, eqn (7.4) may be written

dv dV v v _ - - -kdp +- BdT (7.13)

The ideal gas law applies to gases in the superheat phase. Th~s is because when gases are superheated they obey the kinetic theory of gases in which the following assumptions are made:

0 molecules are solid spheres; 0 molecules occupy a negligible proportion of the total volume of the gas; 0 there are no forces of attraction between the molecules, but there are infinite forces of

repulsion on contact.

If a gas is not superheated the molecules are closer together and the assumptions are less valid. This has led to the development of other models which take into account the interactions between the molecules.

7.2 Van der Waals’ equation of state

In an attempt to overcome the limitations of the perfect gas equation, a number of modifications have been made to it. The two most obvious modifications are to assume:

e the diameters of the molecules are an appreciable fraction of the mean distances between them, and the mean free path between collisions - this basically means that the volume occupied by the molecules is not negligible;

0 the molecules exert forces of attraction which vary with the distance between them, while still exerting infinite forces of repulsion on contact.

First, if it is assumed that the molecules occupy a significant part of the volume occupied by the gas then eqn (7.10) can be modified to

%T P = -

v - b (7.14)

where b is the volume occupied by the molecules. This is called the Clausius equation of state.

Page 140: Advanced Thermodynamics for Engineers

124 Equations of state

Second, allowing for the forces of attraction between molecules gives the van der Waals’ equation of state, which is written

%T a p = - - -

v - b v 2 (7.15)

Figure 7.1 shows five isotherms for water calculated using van der Waals’ equation (the derivation of the constants in the eqn (7.15) is described below). It can be seen that eqn (7.15) is a significant improvement over the perfect gas law as the state of the water approaches the saturated liquid and vapour lines. The line at 374°C is the isotherm at the critical temperature. This line passes through the critical point, and follows closely the saturated vapour line (in fact, it lies just in the two-phase region, which indicates an inaccuracy in the method). At the critical temperature the isotherm exhibits a point of inflection at the critical point. At temperatures above the critical temperature the isotherms exhibit monotonic behaviour, and by 500°C the isotherm is close to a rectangular hyperbola, which would be predicted for a perfect gas. At temperatures below the critical isotherm, the isotherms are no longer monotonic, but exhibit the characteristics of a cubic equation. While this characteristic is not in agreement with empirical experience it does result in the correct form of function in the saturated liquid region - which could never be achieved by a perfect gas law. It is possible to resolve the problem of the correct pressure to use for an isotherm in the two-phase region by considering the Gibbs function, which must remain constant during the evaporation process (see Chapter 1). This results in a constant pressure (horizontal) line which obeys an equilibrium relationship described in Section 7.4. It can also be seen from Fig 7.1 that the behaviour of a substance obeying the van der Waals’ equation of state approaches that of a perfect gas when

0 the temperature is above the critical temperature 0 the pressure is low compared with the critical pressure.

400

350

300 c d

250 e - 2 200 $ v1

2 150

100

50

0

.oo 1 .o 1 .1 1

Specific volume / (m3/kg)

Fig. 7.1 Isotherms on a p - v diagram calculated using van der Waals’ equation

Page 141: Advanced Thermodynamics for Engineers

Law of corresponding states 125

Exercise

Evaluate the isothermal compressibility, k, and coefficient of expansion, j9, of a van der Waals’ gas.

7.3 Law of corresponding states

If it is assumed that all substances obey an equation of the form defined by the van der Waals’ equation of state then all state diagrams will be geometrically similar. This means that if the diagrams were normalised by dividing the properties defining a particular state point ( p, v , T ) by the properties at the critical point ( pc, vc, T,) then all state diagrams based on these reducedproperties will be identical. Thus there would be a general relationship

(7.16) vR = f ( P R , TR)

where vR = v /vc = reduced volume; pR = p / p c = reduced pressure; TR = T/Tc = reduced temperature.

To be able to define the general equation that might represent all substances it is necessary to evaluate the values at the critical point in relation to the other parameters. It has been Seen from Fig 7.1 that the critical point is the boundary between two different forms of behaviour of the van der Waals’ equation. At temperatures above the critical point the curves are monotonic, whereas below the critical point they exhibit maxima and minima. Hence, the critical isotherm has the following characteristics: @p/av>, and (a2p/av2), = 0.

Consider van der Waals’ equation

RT a p = - - -

v - b v 2

This can be differentiated to give the two differentials, giving

RT 2a + -

2RT 6a

and at the critical point

RT, 2a - + - = o 3 (Vc-bI2 vc

2RTc 6a

(vc-bI3 vC -0

4

giving

(7.17)

(7.18)

(7.19)

(7.20)

8a a 27bR’ 27b2

vC=3b; T C = - p c = - (7.21)

Page 142: Advanced Thermodynamics for Engineers

126 Equations of state

Hence

(7.22)

The equation of state for a gas obeying the law of corresponding states can be written

or

(7.23)

(7.24)

Equation (7.24) is the equation of state for a substance Obeying Van der Waals’ equation: it should be noted that it does not explicitly contain Values of a and b. It is possible to obtain a similar solution which omits a and b for any two-parameter state equation, but such a solution has not been found for state equations with more than two parameters. The law of corresponding states based on van der Waals’ equation does not give a very accurate prediction of the properties of substances over their whole range, but it does demonstrate some of the important differences between real substances and perfect gases. Figure 7.1 shows the state diagram for water evaluated using parameters in van der Waals’ equation based on the law of corresponding states. The parameters used for the calculation of Fig 7.1 will now be evaluated.

The values of the relevant parameters defining the critical point of water are

p c = 221.2 bar; v, = 0.00317 m3/kg; Tc = 374.15°C

which gives

v 0.00317

3 3 b=“= = 0.0010567 m3/kg

a = 3pCv,2 = 3 x 221.2 x 0.003172 = 0.0066684 bar m6/kg20.003172 0.0066684barm6/kg2

8 x 221.2 x 0.00317 3 x (374.15 + 273)

R = = 0.002889 bar m3/ kg K

Thus, the van der Waals’ equation for water is

0.0066684 - 0.002889T P =

(v- 0.0010567) V 2 (7.25)

Note that the value for ‘R ’ in this equation is not the same as the specific gas constant for the substance behaving as a perfect gas. This is because it has been evaluated using values of parameters at the critical point, which is not in the region where the substance performs as a perfect gas.

It is possible to manipulate the coefficients of van der Waals’ equation to use the correct value of R for the substance involved. This does not give a very good approximation for the critical isotherm, but is reasonable elsewhere. Considering eqn (7.22), the term for vc

2

c

v

3 3 T

Joe Sulton
Joe Sulton
Page 143: Advanced Thermodynamics for Engineers

can be eliminated to give

Law of corresponding states 127

(7.26)

Substituting the values of p c and Tc into these terms, using R = %/18, gives

27 x (8.3143 x 103/18)2 x (374.1 + 273)2 a =

64 x 221.2 x 10’

= 1703.4Pam6/kg2 = 0.017034 barm6/kg2 8.3143 x lo3 x (374.1 + 273)

18 x 8 x 221.2 x 10’ b = = 0.0016891 m3/kg

which results in the following van der Waals’ equation for water:

0.004619 T 0.0 17034 - P = (V - 0.0016891) V 2

(7.27)

It can be readily seen that eqn (7.27) does not accurately predict the critical isotherm at low specific volumes, because the value of b is too big. However, it gives a reasonable prediction of the saturated vapour region, as will be demonstrated for the isotherms at 200”C, 300°C and the critical isotherm at 374°C. These are shown in Fig 7.2, where the predictions are compared with those from eqn (7.25).

400

350

300 l d

. 8

200 2 ul g 150

100

50

0

- Equation 25

--c Equation 27

Saturated l iquid + vapour lines

.001 .o 1 .1 1

Specif ic volume / (m3/kg)

Fig. 7.2 Comparison of isotherms calculated by van der Waals’ equation based on eqns (7.25) and (7.27)

It is interesting to compare the values calculated using van der Waals’ equation with those in tables. This has been done for the isotherm at 200”C, and a pressure of 15 bar (see Fig 7.3), and the results are shown in Table 7.1.

Page 144: Advanced Thermodynamics for Engineers

128 Equations of state

Table 7.1 Specific volume at 200°C and 15 bar

Specific volume from Specific volume from Specific volume Region eqn (7.25)/ (m3/kg ) eqn (7.27) / (m 3/kg 1 from tables/(m3/kg)

Saturated liquid 0.00 16 0.0017 0.001 157

Saturated vapour 0.09 0.14 0.1324

2 5

20 h c m e 1 15

e, c

2 2 10 (II

PI

5

0

001

I , ' ( ' 1 I

.o 1 .1 1

Specific volume / (m3 /kg)

Fig. 7.3 Expanded diagram of 200°C isotherms, showing specific volumes at 15 bar

The above diagrams and tables show that van der Waals' equation does not give a good overall representation of the behaviour of a gas in the liquid and mixed state regions. However, it is a great improvement on the perfect, or ideal, gas equation in regions away from superheat. It will be shown later that van der Waals' equation is capable of demonstrat- ing certain characteristics of gases in the two-phase region, e.g. liquefaction. There are other more accurate equations for evaluating the properties of substances, and these include

0 the virial equation:

0 the Beattie-Bridgeman equation:

(7.28)

RT A p = 7 (1 - e ) (v + B ) - 7

V V

Page 145: Advanced Thermodynamics for Engineers

Isotherms or isobars in the two-phase region 129

where

A =Ao(i - f),

0 the Bertholet equation:

RT a p=---

v - b Tv2

0 the Dieterici equation:

RT -alRTv p"- e v - b

C B = B o l - - , e = - ( :) vT3

Returning to the van der Waals' gas, examination of eqns (7.22) shows that

z, = - pcvc = - = 0.375 RT, 8

(7.29)

(7.30)

(7.31)

(7.32)

This value z, (sometimes denoted p,) is the compression coefficient, or compressibility factor, of a substance at the critical point. In general, the compression coefficient, z, is defined as

P V RT

z = - (7.33)

It is obvious that, for a real gas, z is not constant, because z = 1 for a perfect gas (i.e. in the superheat region), but it is 0.375 (ideally) at the critical point. There are tables and graphs which show the variation of z with state point, and these can be used to calculate the properties of real gases.

7.4 Isotherms or isobars in the two-phase region

Equation (7.15) predicts that along an isotherm the pressure is related to the specific volume by a cubic equation, and this means that in some regions there are three values of volume which satisfy a particular pressure. Since the critical point has been defined as an inversion point then the multi-valued region lies below the critical point, and experience indicates that this is the two-phase (liquid + vapour) region. It is known that pressure and temperature are not independent variables in this region, and that fluids evaporate at constant temperature in a constant pressure chamber; hence these isotherms should be at constant pressure. This anomaly can be resolved by considering the Gibbs energy of the fluid in the two-phase region. Since the evaporation must be an equilibrium process then it must obey the conditions of equilibrium. Considering the isotherm shown in Fig 7.4, which has been calculated using eqn (7.27) for water at 300°C, it can be seen that in the regions from 3 to 1 and from 7 to 5, a decrease in pressure results in an increase in the specific volume. However, in the region between 5 and 3 an increase in pressure results in an increase in specific volume: this situation is obviously unstable (see also Chapter 6). It was shown in Chapter 1 that equilibrium was defined by dG I p , T L 0. Now

d g = v d p - s d T (7.34)

Page 146: Advanced Thermodynamics for Engineers

130 Equations of state

and hence along an isotherm the variation of Gibbs function from an initial point, say, 1 to another point is

(7.35)

This variation has been calculated for the region between points 1 and 7 in Fig 7.4, and is shown in Fig 7.5.

0 0.005 0.01 0.015 0.02 0.025 0.03 Specific volume / (m3/kg)

Fig. 7.4 Variation of pressure and specific volume for water along the 30O0C isotherm

It can be seen from Fig 7.5 that the Gibbs function increases along the isotherm from point 1 to point 3. Between point 3 and point 5 the Gibbs function decreases, but then begins to increase again as the fluid passes from 5 to 7. The regions between 1 and 3 and 5 and 7 can be seen to be stable in terms of variation of Gibbs function, whereas that between 3 and 5 is unstable: this supports the argument introduced earlier based on the variation of pressure and volume. Figure 7.5 also shows that the values of Gibbs energy and pressure are all equal at the points 2, 4 and 6, which is a line of constant pressure in Fig 7.4 This indicates that the Gibbs function of the liquid and vapour phases can be equal if the two ends of the evaporation process are at equal pressures for an isotherm. Now considering the region between the saturated liquid and saturated vapour lines: the Gibbs function along the ‘isotherm’ shown in Fig 7.4 will increase between points 2 and 3, and the substance would attempt to change spontaneously back from 3 to 2, or from 3 to 4, and hence point 3 is obviously unstable. In a similar manner the stability of point 5 can be considered. An isobar at 100 bar is shown in Fig 7.5, and it can be seen that it crosses the Gibbs function line at three points - the lowest Gibbs function is at state 1. Hence, state 1 is more stable than the other points. If that isobar were moved down further until it passed through point 5 then there would be other points of higher stability than point 5 and the system would tend to move to these points. This argument has still not defined the equilibrium line

Page 147: Advanced Thermodynamics for Engineers

Problems 131

200

175

1 so

100

7s

so

3

5 5

I

Variation in Gibbs function along isotherm

Fig. 7.5 Variation of Gibbs function for water along the 300°C isotherm

between the liquid and vapour lines, but this can be defined by ensuring that the Gibbs function remains constant in the two-phase region, and this means the equilibrium state must be defined by constant temperature and constant Gibbs function. Three points on that line are defined by states 2, 4 and 6. The pressure which enables constant Gibbs function to be achieved is such that

5,4” vdp = 0

and this means that

(7.36)

(7.37) -I

Region I Region 11

which means that the area of the region between the equilibrium isobar and line 2-3-4 (Region I) must be equal to that between the isobar and line 4 - 5 - 6 (Region II ). This is referred to as Marwell’s equal area rule, and the areas are labelled I and II in Fig 7.4.

7.5 Concluding remarks

A number of different equations of state have been introduced which can describe the behaviour of substances over a broader range than the common perfect gas equation. Van der Waal’s equation has been analysed quite extensively, and it has been shown to be capable of defining the behaviour of gases close to the saturated vapour line. The law of corresponding states has been developed, and enables a general equation of state to be considered for all substances. The region between the saturated liquid and vapour lines has been analysed, using Gibbs energy to define the equilibrium state.

Page 148: Advanced Thermodynamics for Engineers

132 Equations of state

PROBLEMS

The Dieterici equation for a pure substance is given by

v - b Determine

(a) the constants a and b in terms of the critical pressure and temperature; (b) the compressibility factor at the critical condition; (c) the law of corresponding states.

Derive expressions for (&,/dv), for substances obeying the following laws:

ST e-(a/%Tv) (a) P = -

v - b

R T a v - b Tv2

(b) p = - - -

U c + -. R T (c) p = - - ___

v - b V ( V - b ) v3

Discuss the physical implication of the results.

The difference of specific heats for an ideal gas, c ~ , ~ - cV,,, = 3. Evaluate the difference in specific heats for gases obeying (a) the van der Waals', and (b) the Dieterici equations of state. Comment on the results for the difference in specific heat for these gases compared with the ideal gas.

Derive an expression for following expression:

R T a p = - - -

v - b T v 2

the law of corresponding states for a gas represented by the

Page 149: Advanced Thermodynamics for Engineers

Problems 133

A, for a gas obeying the state equation

5 p v = ( l + a ) % T

where a is a function of temperature alone, that the specific heat at constant pressure is given by

d2 ( a T )

dT2 cp = -%T ~ h P + CP"

where cpo is the specific heat at unit pressure.

6 The virial equation of state is

Compare this equation with van der Waals' equation of state and determine the first two virial coefficients, b, and b,, as a function temperature and the van der Waals' constants.

Determine the critical temperature and volume (T,, v,) for the van der Waals' gas, and show that

27 T, b - _ 2 - v c 3 (

BV. ) [b, = 1; b2= ( b - u/%T)]

7 The equation of state for a certain gas is

where A is constant.

then the value of the specific heat at constant pressure at the state (T , p ) is given by Show that if the specific heat at constant pressure at some datum pressure p o is cpo

cp- cPo=%TA e-AT(2-AT)(p-po)

8 How can the equation of state in the form of a relationship between pressure, volume and temperature be used to extend limited data on the entropy of a substance.

A certain gas, A, has the equation of state

pv=%T(l+ a p )

where a is a function of temperature alone. Show that

Another gas B behaves as an ideal gas. If the molar entropy of gas A is equal to that of gas B when both are at pressure p o and the same temperature T, show that if the

5 Show

Joe Sulton
Page 150: Advanced Thermodynamics for Engineers

134 Equations of state

pressure is increased to p with the temperature maintained constant at T , tht, entropy of .gas B exceeds that of gas A by an amount

%(p -po) a + T - ( :;) 9 A gas has the equation of state

-- pv - a - b T RT

where u and b are constants. If the gas is compressed reversibly and isothermally at the temperature T' show that the compression will also be adiabatic if

a 2b

T' = -

Page 151: Advanced Thermodynamics for Engineers

Liquefaction of Gases

If the temperature and pressure of a gas can be brought into the region between the satur- ated liquid and saturated vapour lines then the gas will become ‘wet’ and this ‘wetness’ will condense giving a liquid. Most gases existing in the atmosphere are extremely superheated, but are at pressures well below their critical pressures. Critical point data for common gases and some hydrocarbons are given in Table 8.1.

Table 8.1 Critical temperatures and pressures for common substances

Substance Critical temperature Critical pressure, T,OC [KI P c (bar)

374 [647]

32 [305] 96 [369]

153 [426] 31 [304]

-82 [191]

-130 [143] -243 [30] -147 [126]

221.2 46.4 49.4 43.6 36.5 89 51 13 34

Figure 8.1 depicts qualitatively the state point of oxygen at ambient conditions and shows that it is a superheated gas at this pressure and temperature, existing at well above the critical temperature but below the critical pressure. If it is desired to liquefy the gas it is necessary to take its state point into the saturated liquid-saturated vapour region. This can be achieved in a number of ways. First, experience indicates that ‘heat’ has to be taken out of the gas. This can be done by two means:

(i) cooling the gas by heat transfer to a cold reservoir, i.e. refrigeration; (ii) expanding the gas in a reversible manner, so that it does work.

8.1

This method is satisfactory if the liquefaction process does not require very low temperatures. A number of common gases can be obtained in liquid form by cooling.

Liquefaction by cooling - method (i)

Page 152: Advanced Thermodynamics for Engineers

136 Liquefaction of gases

Fig. 8.1

Entropy, S

State point of oxygen at ambient temperature and pressure

Examples of these are the hydrocarbons butane and propane, which can both exist as liquids at room temperature if they are contained at elevated pressures. Mixtures of hydrocarbons can also be obtained as liquids and these include liquefied petroleum gas (LPG) and liquefied natural gas (LNG).

A simple refrigerator and the refrigeration cycle are shown in Figs 8.2(a) and (b) respectively.

4

k

Condenser

4

k Throttle

Evaporator I d *

b

gj

k! E

c”

I

I I Q i n Entropy, S

Fig. 8.2 A simple refrigerator: (a) schematic of plant; (b) temperature-entropy diagram

Consider the throttling process in Fig 8.2, which is between 4 and 1. The working fluid enters the throttle at a high pressure in a liquid state, and leaves it at a lower pressure and temperature as a wet vapour. If the mass of working fluid entering the throttle is 1 kg then the mass of liquid leaving the throttle is (1 - x , ) kg. If this liquid were then withdrawn to a vessel, and the mass of fluid in the system were made up by adding (1 - x, ) kg of gas at

Page 153: Advanced Thermodynamics for Engineers

Liquefaction by cooling - method ( i ) 137

state 3 then it would be possible to liquefy that gas. The liquefaction has effectively taken place because, in passing through the throttle, the quality (dryness fraction) of the fluid has been increased, and the energy to form the vapour phase has been obtained from the latent heat of the liquid thus formed: the throttling process is an isenthalpic one, and hence energy has been conserved. If the working fluid for this cycle was a gas, which it was desired to liquefy, then the liquid could be withdrawn at state 1’.

The simple refrigeration system shown in Fig 8.2 might be used to liquefy substances which boil at close to the ambient temperature, but it is more common to have to use refrigeration plants in series, referred to as a cascade, to achieve reasonable levels of cooling. The cascade plant is also more cost-effective because it splits the temperature drop between two working fluids. Consider Fig 8.2: the difference between the top and bottom temperatures is related to the difference between the pressures. To achieve a low temperature it is necessary to reduce the evaporator pressure, and this will increase the specific volume of the working fluid and hence the size of the evaporator. The large temperature difference will also decrease the coefficient of performance of the plant. The use of two plants in cascade enables the working fluid in each section to be optimised for the temperature range encountered. Two plants in cascade are shown in Fig 8.3 (a), and the T-s diagrams are depicted in Fig 8.3(b).

2

I

I Entropy, S

(b)

Fig. 8.3 Cascade refrigeration cycle

In the cascade arrangement the cooling process takes place in two stages: it is referred to as a binary cascade cycle. This arrangement is the refrigeration equivalent of the combined cycle power station. The substance to be liquefied follows cycle 1-2-3-4-1, and the liquid is taken out at state 1’. However, instead of transferring its waste energy, Q,, to the environment it transfers it to another refrigeration cycle which operates at a higher temperature level. The working fluids will be different in each cycle, and that in the high temperature cycle, 5-6-7-8-5, will have a higher boiling point than the substance being

Page 154: Advanced Thermodynamics for Engineers

138 Liquefaction of gases

liquefied. The two cycles can also be used to liquefy both of the working fluids, h. case liquid will also be taken out at state 5’.

evaluated by considering the heat flow through each section and is given by The overall coefficient of performance of the two plants working in cascade can be

(1 + ;)=I +;)(I + ;) (8.1)

where B’is the overall coefficient of the combined plant, and B f l and /?’* are the coefficients of performance of the separate parts of the cascade. The coefficient of performance of the overall plant is less than that of either individual part.

The liquefaction of natural gas is achieved by means of a ternary cascade refrigeration cycle, and the components are shown in Fig 8.4. In this system a range of hydrocarbons are used to cool each other. The longer chain hydrocarbons have higher boiling points than the shorter ones, and hence the alkanes can be used as the working fluids in the plant.

Fig. 8.4 Plant for liquefaction of natural gas

Page 155: Advanced Thermodynamics for Engineers

Liquefaction by cooling - method ( i ) 139

Another common gas that can be obtained in a non-gaseous form is carbon dioxide (CO,). This is usually supplied nowadays in a solid form called dry ice, although it was originally provided as a pressurised liquid. Dry ice is obtained by a modification of the refrigeration process, and the plant and relevant T-s diagrams are shown in Figs 8.5(a) and (b) respectively. This is like a ternary cascade system, but in this case the working fluid passes through all the stages of the plant. The carbon dioxide gas enters the plant at state ‘a’, and passes through the final stage, referred to as the snow chamber, because by this stage the gas has been converted to dry ice. Heat is transferred from the carbon dioxide to the final product causing some evaporation: this evaporated product is passed back to the first compressor. The fluid then passes to the first compressor, where it is compressed to state ‘c’. It is then passed into a flash tank where it is cooled by transferring heat with liquid already in there; this again causes some evaporation of the product, which is passed back into the second compressor, along with the working fluid. In the example shown the process has three stages, and the final stage must be at a pressure high enough

Fig. 8.5 Plant for manufacture of dry ice (solid carbon dioxide)

Page 156: Advanced Thermodynamics for Engineers

140 Liquefaction of gases

temperature

m

‘,. Entropy, S

Fig. 8.5 Continued

to enable the condensation to occur by heat transfer to a fluid at close to atmospheric temperature, depicted by point ‘h’ in Fig 8.5(b); the pressure must also be below the critical pressure for condensation to occur. The liquid that is produced at ‘j’ passes through the series of throttles to obtain the refrigeration effect necessary to liquefy the carbon dioxide. In the final chamber the liquid is expanded below the triple point pressure, and the phase change enters the solid-gas region. Carbon dioxide cannot exist as a liquid at atmospheric pressure, because its triple point pressure is 5.17 bar; hence it can be either a pressurised liquid or a solid. Dry ice is not in equilibrium with its surroundings, and is continually evaporating. It could only be in a stable state if the ambient temperature were dropped to the equivalent of the saturated temperature for the solid state at a pressure of 1 bar.

8.2 Liquefaction by expansion - method (ii)

If the gas does work against a device (e.g. a turbine) whilst expanding adiabatically then the internal energy will be reduced and liquefaction may ensue. A cycle which includes such an expansion process is shown in Fig 8.6. The gas, for example oxygen, exists at state 1 when at ambient temperature and pressure: it is in a superheated state, but below the critical pressure. If the gas is then compressed isentropically to a pressure above the critical pressure it will reach state 2. The temperature at state 2 is above the ambient temperature, T,, but heat transfer to the surroundings allows the temperature to be reduced to state point 2a. If the liquefaction plant is a continuously operating plant, then there will be available a supply of extremely cold gas or liquid and this can be used, by a suitable arrangement of heat exchangers, to cool the gas further to state point 3. If the gas is now expanded isentropically through the device from state 3 down to its original pressure it will condense out as a liquid at state 4. Hence, the processes defined in Fig 8.6 can be used to achieve liquefaction of a gas by use of a device taking energy out of the substance by producing work output. It should be noted that if the gas was a supercritical vapour at ambient conditions it would be impossible to obtain it in liquid form at atmospheric pressure.

Page 157: Advanced Thermodynamics for Engineers

The Joule-Thomson eaect 141

In the liquefaction process described above, the gas may be expanded from state 3 to state4 either by a reciprocating machine or against an expansion turbine. Both these machines suffer the problem that the work done in the expansion process is a function of the initial temperature because the temperature drop across the turbine is

It can be seen from eqn (8.2) that as T3 becomes very small, the temperature drop achieved by the expansion gets smaller, which means that the pressure ratio to obtain the same temperature drop has to be increased for very low critical temperatures. Another major problem that occurs at very low temperatures is that lubrication becomes extremely difficult. For this reason the turbine is a better alternative than a reciprocating device because it may have air bearings, or gas bearings of the same substance as that being liquefied, thus reducing contamination.

2

,’ p = l bar

1

Entropy, S

Fig. 8.6 A method for liquefying a gas

Fortunately another method of liquefaction is available which overcomes many of the problems described above. This is known as the Joule-Thomson effect and it can be evaluated analytically. The Joule-Thomson effect is the result of relationships between the properties of the gas in question.

8.3 The Joule-Thomson effect

The Joule-Thomson effect was discovered in the mid-19th century when experiments were being undertaken to d e h e the First Law of Thermodynamics. Joule had shown that the specific heat at constant volume was not a function of volume, and a similar experiment was developed to ascertain the change of enthalpy with pressure. The

Page 158: Advanced Thermodynamics for Engineers

142 Liqwfaction of gases

experiment consisted of forcing a gas through a porous plug by means of a pressure drop. It was found that, for some gases, at a certain entry temperature there was a temperature drop in the gas after it had passed through the plug. This showed that, for these gases, the enthalpy of the gas was a function of both temperature and pressure. A suitable apparatus for conducting the experiment is shown in Fig 8.7.

b

g G z

Piston maintaining - - - - - -

- - - - - - I

Piston maintaining

Isenthalpic _Ti_ - - - - - - - - - - - - - - - - -

TI /y - - - - - - - - . . . . . . . . . . . . . . . . . . . . I I I I I I I I I I I I I I I I

I I I I I I I I I I I I I I I I I Inversion I

pressure IPi ' PI

C0;ltrol Porous surface Plug

~~~~~ ~ ~ ~

Fig. 8.7 Porous plug device for the Joule-Thomson experiment

Page 159: Advanced Thermodynamics for Engineers

The Joule-Thomson effect 143

The process shown in Fig 8.8 may be analysed by applying the steady flow energy equation (SFEE) across control volume, when

where

0 = rate of heat transfer, W = rate of work transfer, lit = mass flow rate, h,, h2 = enthalpy upstream and downstream of plug, V, , V, = velocity upstream and downstream of plug.

Now Q = 0, I$' = 0, V2 = V,. Hence h2 = hl , and the enthalpy along line ABC is constant; thus ABC is an isenthulpic line. The most efficient situation from the viewpoint of obtaining cooling is achieved if p1 and T, are at B, the inversion point. If state ( pl, T,) is to the right of B then T2 depends on the pressure drop, and it could be greater than, equal to or less than T,. If the upstream state ( p , , T , ) is at B then the downstream temperature, T2, will always be less than TI. It is possible to analyse whether heating or cooling will occur by evaluating the sign of the derivative (aT/ap),. This term is called the Joule-Thomson coefficient, p.

If p<O, then there can be either heating or cooling, depending on whether the downstream pressure is between A and C, or to the left of C. If p > 0, then the gas will be cooled on passing through the plug (i.e. the upstream state point defined by ( p , , T , ) is to the left of B on the isenthalpic line).

This situation can be analysed in the following way. From the Second Law of Thermodynamics

d h = T ds+vdp (8.4)

But, for this process dh = 0, and thus

o-I($) h + v

If it is assumed that entropy is a continuous function of pressure and temperature, i.e. s = s( p, T), then

ds = ( $)T dp + ($):T

which can be rearranged to give

($)h = ($)T + ( $)p ($), Hence, substituting this expression into eqn (8.5) gives

(8.6)

(8.7)

Page 160: Advanced Thermodynamics for Engineers

1 4 4 Liquefaction of gases

Now, from the thermodynamic relationships (eqn (6.24)),

and, from the Maxwell relationships (eqn <6.19d)),

Thus

which may be rearranged to give

(2). = ; [.( g)p - V I = P

This can be written in terms of the coefficient of expansion

and eqn (8.10) becomes V

p = - [ B T - l ] CP

At the inversion temperature,

($).=(I T i = - 1

B

(8.9)

(8.10)

(8.1 1)

(8.12)

It is possible to evaluate the Joule-Thomson coefficient for various gases from their state equations. For example, the Joule-Thomson coefficient for a perfect gas can be evaluated by evaluating ,u from eqn (8.10) by differentiating the perfect gas equation

pv = RT (8.13) Now

and, from eqn (8.13), the derivative

Hence

(8.14)

Page 161: Advanced Thermodynamics for Engineers

The Joule-Thomson effect 145

This means that it is not possible to cool a perfect gas by the Joule-Thomson effect. This is what would be expected, because the enthalpy of an ideal gas is not a function of pressure. However, it does not mean that gases which obey the ideal gas law at normal atmospheric conditions (e.g. oxygen, nitrogen, etc) cannot be liquefied using the Joule-Thomson effect, because they cease to obey this law close to the saturated vapour line. The possibility of liquefying a gas obeying the van der Waals' equation is considered below.

Van der Waals' equation for air may be written, using a and b for air as 1.358 bar (m3/ km01)~ and 0.0364 m3/kmol respectively, as

0.083143 T 1.358 -- P' (V-0 .0364) V'

(8.15)

To assess whether a van der Waals' gas can be liquefied using the Joule-Thomson effect it is necessary to evaluate the Joule-Thomson coefficient, p. This is related to (av /aT) , , which can be evaluated in the following way.

From eqn (8.15), in general form

and hence,

v - b 2a

= f (p + 5) - (7) (7) giving

l a v \ %

(8.16)

(8.17)

Thus, from eqn (8.17), the Joule-Thomson coefficient for a van der Waals' gas is not zero at all points, being given by

(8.18)

8.3.1 MAXIMUM AND MINIMUM IWERSION TEMPERATURES

The general form of van der Waals' equation, eqn (8.15), can be rewritten as

The inversion temperature is defined as

Ti =(E) P

(8.19)

Page 162: Advanced Thermodynamics for Engineers

146 Liquefaction of gases

which can be evaluated from eqn (8.16) in the following way:

v - b ) 2a ($)p = ; (P + $) - (b) (7)

* v($)p= + $) - (v - b ) ( 3

(8.20)

It is now necessary to solve for v in terms of multiplying q n (8.16) by v and eqn (8.20) by ( v - b). This gives

alone, and this can be achieved by

and

(v - b)Ti = e ( p + ;) - - ( v - b)2 - 2a R R v2

Subtracting eqns (8.21) and (8.22) gives

(I+/%- = x = l - - b V

Substituting eqn (8.23) in eqn (8.15) gives

(8.21)

(8.22)

(8.23)

(8.24)

The maximum and minimum inversion temperatures are achieved when the pressure, p, is zero. This gives

2a - 2a Ti Rb 9Rb 9

pi= - and T i = - = -

The value of the critical temperature for a van der Waals' gas is

8a Tc = -

27Rb

and hence

Ti Maximum inversion temperature Tc Critical temperature

Ti Minimum inversion temperature TC Critical temperature

= 6.75 -- -

= 0.75 -- -

(8.25)

(8.26)

(8.27a, b)

These equations will now be applied to air. The values of a and b for air are 1.358 bar (m3/kmol)2 and 0.0364 m3/kmol respectively, and van der Waals' equation for air was given as eqn (8.15). Substituting these values into eqn (8.24), which can be solved as a

Page 163: Advanced Thermodynamics for Engineers

The Joule-Thomon effect 147

quadratic equation in K , gives the diagram shown in Fig 8.9. This shows that it is not possible to liquefy air if it is above a pressure of about 340 bar.

0 e 50 100 150 w 250 MO 3 9

Pressure I (bar)

Fig. 8.9 Higher and lower inversion temperatures for air

The maximum pressure for which inversion can occur is when dp/dT = 0. This can be related to van der Waals’ equation in the following way, using the expression for x introduced in eqn (8.23):

dp dp dx dT dx dT -- ---

and since dx/dT is a simple single term expression

dP dP - = O when - = 0 dT dx

-= - dp dx b2

a {3(1 -x) - (3x - 1))

Equation (8.28) is zero when 3(1- x) - (3x - 1) = 0, Le. when x = 2/3. Hence the maximum pressure at which inversion can occur is

2

(8.28)

(8.29)

Thus the maximum pressure at which a Van der Waals’ gas can be liquefied using the Joule-Thomson effect is nine times the critical pressure. Substituting for a and b in eqn (8.29) gives the maximum pressure as 341.6 bar.

Figure 8.9 indicates that it is not possible to cool air using the Joule-Thomson effect if it is at a temperature of greater than 900 K, or less than about 100 K. Similar calculations for hydrogen give 224 K and 24.9 K, respectively, for the maximum and minimum inversion temperatures. The maximum pressure at which inversion can be achieved for hydrogen is 117 bar.

Page 164: Advanced Thermodynamics for Engineers

148 Liquefaction of gases

The inversion curve goes through the lines of constant enthalpy at the pow. ( a T / a p ) , = 0. To the left of the inversion curve, cooling always occurs; to the %Le heating or cooling occurs depending on the values of the pressure.

Figure 8.9 may be drawn in a more general manner by replotting it in terms of the reduced pressure ( pR) and temperature (TR) used in the law of corresponding states. This generalised diagram is shown in Fig 8.10. Also shown on that diagram is the saturation line for water plotted in non-dimensional form. For a gas to be cooled using the Joule-Thomson effect it is necessary that its state lies in the region shown in Fig 8.10. Consideration of Table 8.1 shows that for oxygen and nitrogen at atmospheric temperature the values of reduced temperature are 2.10 and 2.38 respectively. Hence, these gases can both be liquefied at pressures up to 8p,, i.e. approximately 400 and 270 bar respectively. However, hydrogen at atmospheric temperature has a TR value of 10, and lies outside the range of the Joule-Thomson effect. To be able to liquefy hydrogen by this method it is necessary to precool it by using another fluid, e.g. liquid nitrogen when the value of TR could become as low as 4.2, which is well within the range of inversion temperatures.

- lo l p ~ p - ~ u l e - T h o m s o n cooling cannot occur in this region1

8

01

3 % 6

n

" 4

L

B

2

0

Joule-Thomon coeficient negative in

_ ~ _ _ - ~

Joule-Thomson coefficient - positive in this region

0 1 2 3 4 5 6 1

Reduced temperature

Fig. 8.10 Inversion region in terms of reduced pressure ( p R = p / p , ) and reduced temperature (T, = T/T,). Note the saturation line

8.4 Linde liquefaction plant

Many gases are liquefied using the Joule-Thomson effect. This approach is embodied in the Linde liquefaction process, and a schematic of the equipment is shown in Fig 8.1 1 (a), while the thermodynamic processes are depicted in Fig 8.11(b). The Linde process is similar to a refrigerator operating on a vapour compression cycle (i.e. the typical refrigeration cycle), except that it includes a heat exchanger to transfer energy equal to QA

and QB from the high temperature working fluid to that which has already been cooled through throttling processes. The two throttling processes depicted in Fig 8.1 1 (a) both bring about cooling through the Joule-Thomson effect. The operating processes in the Linde plant will now be described; it will be assumed that the plant is already operating in steady state and that a supply of liquefied gas exists. The gas is supplied to a compressor at state 1 and make-up gas is supplied at state 11, which is the same as state 1. This is then

Page 165: Advanced Thermodynamics for Engineers

Linde liquefaction plant 149

Compressor

Throttle '8 6

(a)

5- 6 a E

I

Electric motor

' Y Make-up

I1 gas

Y0 - Y

10

1

(b) Entropy, S

Fig. 8.11 Linde liquefaction plant: (a) schematic diagram; (b) T-s diagram

compressed to a high, supercritical pressure (which might be hundreds of atmospheres) by a multi-stage reciprocating compressor with inter-stage cooling. The gas finally reaches state 2, and it is then passed through a heat exchanger which cools it to state 3. At this point it is throttled for the first time, and is cooled by the Joule-Thomson effect to state 4 and passed into a receiver. Some of the gas entering the receiver is passed back to the

Page 166: Advanced Thermodynamics for Engineers

150 Liquefaction of gases

compressor via the heat exchanger, and the remainder is passed through a second throttle until it achieves state 6. At this stage it is in the liquid-vapour region of the state diagram, and liquid gas can be removed at state 9. The yield of liquid gas is y, defined by the quality of state point 6 on the T-s diagram.

8.5 Inversion point on p-v-Tsurface for water The Joule-Thomson effect for water was met early in the study of thermodynamics. It was shown that the quality (dryness fraction) of water vapour could be increased by passing the wet vapour through a throttle. This process is an isenthalpic one in which the pressure decreases. The liquid-vapour boundary for water is shown in Fig 8.12, and it can be seen that the enthalpy of the vapour peaks at a value of about 2800 M/kg. If water vapour at 20 bar, with an enthalpy of 2700 M/kg, is expanded through a throttle then its quality increases; this means that it would not be possible to liquefy the water vapour by the Joule-Thomson effect, and hence this point must be below the minimum inversion temperature. Using the van der Waals' relationship and substitut- ing for the critical temperature gives fi = 0.75 x T, = 212OC. This is equivalent to a pressure of about 20 bar, which is close to that defining the maximum enthalpy on the saturation line. This point of minimum inversion temperature can also be seen on the Mollier ( h - s ) diagram for steam, and is where the saturation curve peaks in enthalpy value. If the pressure were, say, 300 bar and the enthalpy were 2700 M/kg then an isenthalpic expansion would result in the gas becoming wetter, and liquefaction could occur.

3000

9 2500

3. 2000

h a - 2 1500

8 0 1000 % & 500

*

0

v)

0

Fig. 8.12 Variation of enthalpy on the saturated liquid and vapour lines with pressure for steam

Examp le

A simple Linde liquefaction plant is shown in Fig 8.13. This is similar to that in Fig 8.11 except that the liquefaction takes place in a single process, and the cascade is omitted. The plant is used to liquefy air, which is fed to the compressor at 1 bar and 17"C, and compressed isothermally to 200 bar. The compressed air then transfers heat in a counterflow heat exchanger, which has no external heat losses or friction, with the stream leaving the flash chamber, and the inlet temperature of the hot stream equals the outlet

Page 167: Advanced Thermodynamics for Engineers

Inversion point on p -v-T sudace for water 15 1

Liquid yield

290 K ' I 7

(b) Entropy, S

Fig. 8.13 Simplified Linde gas liquefaction plant: (a) schematic diagram of simplified plant; (b) T - s diagram for liquefaction process

temperature of the cold stream. Using the table of properties for air at low temperatures and high pressures (Table 8.2), and taking the dead state conditions for exergy as 1 bar and 290 K, evaluate the:

(i) yield of liquid air per kg of compressed fluid; (ii) temperature of the compressed air before the Joule-Thomson expansion process; (iii) minimum work required per kg of liquid air; (iv) actual work required per kg of liquid air; (v) rational efficiency of the plant, and the irreversibilities introduced by the heat

exchanger and the throttle. (example based on Haywood (1980)).

Page 168: Advanced Thermodynamics for Engineers

Table 8.2 Properties of air (from Haywood, 1972)

Enthalpy (W/kmol) Entropy (M/kmol K) Pressure Temperature ( a m > T,(K) hf h, Sf s,

1 81.7 0 5942 0.00 74.00 5 98.5 917 6249 10.33 65.09

10 108.1 1537 6284 16.24 60.55 20 119.8 2506 6127 24.51 54.93 30 127.8 3381 5760 31.17 49.87 35 131.1 3884 5433 35.08 46.92 37.66 (T,) 132.5 4758 4758 41.39 41.39

Properties of air on the saturated liquid and saturated vapour lines

Pressure (atmospheres) Temperature (K) 1 30 40 50 200 400

90

100

110

120

125

130

135

140

145

150

160

170

180

190

200

250

290

300

h

h

h

h

h

h

h

h

h

h

h

h

h

h

h

h

h

h

S

S

S

S

S

S

S

S

S

S

S

S

S

S

S

S

S

S

6190 76.9 6493 80.09 6795 82.97 7092 85.53 7243 86.78 7393 87.96 754 1 89.09 7691 90.16 7842 91.20 7993 92.28 8290 94.2 8582 95.95 8877 97.62 9172 99.21 9464 100.7 10925 107.3 12097 11 1.63 12364 112.6

2343 22.85 2994 28.14 5976 5 1.54 6355 54.44 6664 56.66 6928 58.52 7168 60.14 7602 62.94 7995 65.31 8365 67.42 8718 69.33 9056 71.07 10671 78.27 11912 82.79 12212 83.91

2187 21.38 2719 25.69 3405 31.07 4816 4 1.74 6048 50.68 6468 53.63 6786 55.77 7309 59.15 7761 61.90 8168 64.24 8546 66.16 8909 68.02 10581 75.61 11848 80.2 12154 81.34

21 14 20.48 2595 24.41 3159 28.81 3847 33.99 4698 40.19 5597 46.54 623 1 50.85 6974 55.65 7509 58.89 7966 61.52 8380 63.75 8761 65.69 10491 73.44 11784 78.14 12094 79.28

2086 15.3 2428 18.11 2770 20.8 3111 23.38 345 1 25.86 3789 28.24 4123 30.51 4779 34.75 5409 38.57 6006 41.98 6583 45.21 7116 47.93 9427 58.28 1 1025 64.09 11405 65.52

4944 29.98 5439 32.98 5915 35.87 6390 38.44 6858 40.84 9074 50.73 10738 56.75 11133 58.23

Properties of superheated air at low temperatures and high pressures h: kJ/kmol s: kJ/kmol K

m, = 28.9

Page 169: Advanced Thermodynamics for Engineers

Inversion point on p-v-T surface for water 153

Solution

(i) Yield of liquid per kg of compressed fluid.

At point 1, p 1 = 1 bar; TI = 290 K; hl = 12097 k.J/kmol

at point 2 , p , = 200 bar; T2 = 290 K ; h, = 11025 kJ/kmol sI = 111.63 kJ/kmolK

S, = 64.09 kJ/kmolK

Considering the heat exchanger: since it is adiabatic

H2 - H3 = H7 - H6

Substituting for specific enthalpies in eqn (8.30) gives

m2(h2 - h 3 ) = m6(h7 - h 6 )

Now

m6=m2(1 - y )

and hence

(8.30)

(8.31)

(8.32)

(8.33)

The process from 3 to 4 is isenthalpic, and is a Joule-Thomson process, thus

h4= h3

But

(8.34)

h4=x4hg+ ( 1 - ~ 4 ) h f = ( 1 -y)h6+yh5 (8.35)

Combining eqns (8.33) and (8.34) gives

h2 - h7 h7 - h5

12097 - 11025 12097 - 0

y=-- - = 0.08862

Hence, the yield of liquid air per kg of compressed air is 0.08862 kg.

(ii) The temperature before the Joule-Thomson process. This is the temperature of the gas at point 3. From eqns (8.34) and (8.35)

h3=h4=X4hg+ (1-x4)h f=(1-y)h6+yh5 = ( 1 - 0.08862) x 5942 = 54 15 kJ/kmol

This value of enthalpy at 200 bar is equivalent to a temperature of 170 K.

(8.36)

(8.37)

(iii) Minimum work required per kg liquid yield. Consideration of the control system in Fig 8.13 shows that only three parameters cross the system boundary; these are the make-up gas, the liquid yield, and the work input to the compressor. Hence, the minimum work

Page 170: Advanced Thermodynamics for Engineers

154 Liquefaction of gases

required to achieve liquefaction of the gas is

WWl = Y (bl - bs),

giving the work per unit mass of liquid as

W",, = - W W l = bl - bS (8.38)

In this case, b,=O because point 5 (the liquid point at 1 bar) was chosen as the datum for properties; hence

Y

Gmt = hl - Tosl - 0 = 12097 - 290 x 11 1.63 = -20275 kJ/kmol liquid

(iv) Work required per liquid yield. The compression process is an isothermal one, and hence the work done is

v2 P2 P2 w = p , v , l n - = - p l v l l n - = -RT,ln- V 1 P1 PI

(8.39)

Substituting the values gives

P2 200 PI 1

w = - S T , In- = -8.3143 x 290 x ln - = -12775kJ/kmolgas

- 12775 - 442 - - = -442kJ/kggas = = -4987 kJ /kg liquid

28.9 0.08863

(v) Rational efliciency. The rational efficiency of the plant is defined as

- 701 - 0.1406 %=-=--

%et

w -4987 (8.40)

Hence, the plant is only 14.1% as efficient as it could be if all the energy transfers were reversible. It is instructive to examine where the irreversibilities occur. The heat exchanger The irreversibility of the heat exchanger is

1, = (1 - Y)(b7 - b6) - (b2 - b3) ~0.91137 x [ -20 275 - (-15 518)} - [ -7561 - (-5776)} = -2550 kJ/kmol gas

This is equivalent to 20.0% of the work required to compress the gas. The throttle to achieve liquefaction The irreversibility of the throttle is

IhNe = b, - b3 = 0.91137 x (-15 518) - (-5776) = -8366 H/kmol gas This is equivalent to 65.5% of the work used to compress the gas. Hence, the work required to liquefy the gas is

w = + I , + ZthroDle

Page 171: Advanced Thermodynamics for Engineers

Problems 155

8.6 Concluding remarks

It has been shown that gases can be liquefied in a number of ways. Gases which are liquids at temperatures close to ambient can be liquefied by cooling in a simple refrigeration system. Carbon dioxide, which cannot be maintained as a liquid at ambient pressure, is made into dry ice which is not in equilibrium at room temperature and pressure.

If it is necessary to achieve extremely low temperatures to bring about liquefaction, the Joule-Thomson effect is employed. It is possible to analyse such liquefaction plant using equilibrium thermodynamics and suitable equations of state. The efficiency of liquefaction plant has been calculated and the major influences of irreversibilities in the processes have been illustrated.

PROBLEMS

1 Show that the Joule-Thomson coefficient, p, is given by

.=-(.($),-v) 1

CP

Hence or otherwise show that the inversion temperature (Ti) is

The equation of state for air may be represented by

1.368 -- 9tT P = 2 v,-0.0367 V ,

where p = pressure (bar), T = temperature (K), and v, = molar volume (m3/kmol). Determine the maximum and minimum inversion temperatures and the maximum inversion pressure for air.

[896 K; 99.6 K; 339 bar]

2 The last stage of a liquefaction process is shown in diagrammatic form in Fig P8.2. Derive the relationship between p, and TI for the maximum yield of liquid at conditions pL, TL, hL for a gas obeying the state equation

(p + 7) (v, - 0.0367) = 9tT

where p is pressure (bar), v, is the molar volume (m3/kmol), and T is temperature (K).

Calculate the pressure for maximum yield at a temperature of 120 K.

[p = (1 - 4%) b/z - 1); 62.8 bar]

Page 172: Advanced Thermodynamics for Engineers

156 Liquefaction of gases

I 1 Receiver k = ] - -

- -

Liquefied Y m p a s

Fig. PS.2

3 The equation of state for a certain gas is

R T k v, = - +-

P R T where k is a constant. Show that the variation of temperature with pressure for an isenthalpic process from 1 to 2 is given by

If the initial and final pressures are 50 bar and 2 bar respectively and the initial temperature is 300 K, calculate

(a) the value of the Joule-Thomson coefficient at the initial state, and (b) the final temperature of the gas, given that

k = - 11 .O kJ m3/ (km01)~ cp,,, = 29.0 kJ/kmol K.

4 A gas has the equation of state

-- P V m - 1 + Ap(T3 - 9.75TC T 2 + 9T: T ) + Bp’T R T

[3.041 x lo-’ m3K/J; 298.5 K]

where A and B are positive constants and T, is the critical temperature. Determine the maximum and minimum inversion temperatures, expressed as a multiple of T,.

[6T,; OST,]

Page 173: Advanced Thermodynamics for Engineers

Problems 157

5 A gas has the equation of state

-- pvm - 1 +Np+Mp2 R T

where N and M are functions of temperature. Show that the equation of the inversion curve is

P=-- - dT dNldM dT

If the inversion curve is parabolic and of the form

(T - To), = 4 4 p0 - P ) where To, po and Q are constants, and if the maximum inversion temperature is five times the minimum inversion temperature, show that Q = T:/9p0 and give possible expressions for N and M.

6 In a simple Linde gas liquefaction plant (see Fig 8.13) air is taken in at the ambient conditions of 1 bar and 300 K. The water-jacketed compressor delivers the air at 200 bar and 300 K and has an isothermal efficiency of 70%. There is zero temperature difference at the warm end of the regenerative heat exchanger (i.e. T, = T,). Saturated liquid air is delivered at a pressure of 1 bar. Heat leakages into the plant and pressure drops in the heat exchanger and piping can be neglected.

Calculate the yield of liquid air per unit mass of air compressed, the work input per kilogram if air is liquefied, and the rational efficiency of the liquefaction process.

[7.76%; 8.39 MJ/kg; 8.84%]

Page 174: Advanced Thermodynamics for Engineers

Thermodynamic Properties of Ideal Gases and Ideal Gas Mixtures of

Constant Corn position

It has been shown that it is necessary to use quite sophisticated equations of state to define the properties of vapours which are close the saturated vapour line. However, for gases in the superheat region, the ideal gas equation gives sufficient accuracy for most purposes. The equation of state for an ideal gas, in terms of mass, is

pV=mRT (9.1)

where

p = pressure (N/m2) v = volume (m3> m = mass (kg) R = specific gas constant (kJ/kg K) T = absolute (or thermodynamic) temperature (K)

This can be written in more general terms using the amount of substance, when

p V = n9lT (9.2) where

n = amount of substance, or chemical amount m o l ) 9l = universal gas constant (H/kmol K)

It is found that eqn (9.2) is more useful than eqn (9.1) for combustion calculations because the combustion process takes place on a molar basis. To be able to work in a molar basis it is necessary to know the molecular weights (or relative molecular masses) of the elements and compounds involved in a reaction.

9.1 Molecular weights n e molecular weight (or relative molecular mass) of a substance is the mass of its molecules relative to that of other molecules. The datum for molecular weights is carbn- 12, and this is given a molecular weight of 12. All other elements and compounds have molecular weights relative to this, and their molecular weights are not integers. To be able to perform combustion calculations it is necessary to know the atomic or molecular

Page 175: Advanced Thermodynamics for Engineers

State equation for ideal gases 159

weights of commonly encountered elements; these can be combined to give other compounds. Table 9.1 gives the data for individual elements or compounds in integral numbers (except air which is a mixture of gases); in reality only carbon-12 (used as the basis for atomic/molecular weights) has an integral value, but most values are very close to integral ones and will be quoted as such.

Table 9.1 Molecular weights of elements and compounds commonly encountered in combustion

-

Atmospheric Air 0, N, N, H2 CO CO, H,O C

m, 28.97 32 28 28.17 2 28 44 18 12

9.1 .I AIR As stated previously, most combustion takes place between a hydrocarbon fuel and air. Air is a mixture of gases, the most abundant being oxygen and nitrogen with small proportions of other gases. In fact, in practice, air is a mixture of all elements and compounds because everything will evaporate in air until the partial pressure of its atoms, or molecules, achieves its saturated vapour pressure. In reality, this evaporation can usually be neglected, except in the case of water. Table 9.1 shows that the molecular weight of atmospheric nitrogen is higher than that of pure nitrogen. This is because ‘atmospheric nitrogen’ is taken to be a mixture of nitrogen and about 1.8% by mass of argon, carbon dioxide and other gases; the molecular weight of atmospheric nitrogen includes the effect of the other substances.

The composition of air is defined as 21 % 0, and 79% N, by volume (this can be written 21 mol% 0, and 79 mol% N,). This is equivalent to 23.2% 0, and 76.8% N, by mass.

9.2 State equation for ideal gases

The equation of state for an ideal gas, a , is

p V = mJoT (9.3)

If the mass of gas, ma, is made equal to the molecular weight of the gas in the appropriate units, then the amount of substance a is known as a mol (if the mass is in kg then the amount of substance is called a kmol). If the volume occupied by that this amount of substance is denoted v, then

where v, is the molar specific volume, and has the units of m3/mol or m3/kmol. Now Avogadro’s hypothesis states that

equal volumes of all ideal gases at a particular temperature and pressure contain the same number of molecules (and hence the same amount of substance).

Page 176: Advanced Thermodynamics for Engineers

160 Thermodynamic properties of ideal gases and ideal gas mixtures

Hence, any other gas, b, at the same pressure and temperature will occupy the same volume as gas a , i.e.

If a system contains an amount of substance n, of gas a, then eqn (9.5) may be written

p V , = n,mWoR,T = n,%T (9.6)

It can be seen that for ideal gases the product m,, Ri is the same for all gases: it is called the universal gas constant, 3. The values of the universal gas constant, '%, together with its various units are shown in Table 9.2.

Table 9.2 Values of the universal gas constant, 3, in various units

8314 J/kmol K 1.985 kcal/kmol K 1.985 Btu/lb-mol K 1.985 CHU/lb-mol R 2782 ft lb,/lb-mol K 1545 ft lb,/lb-mol R

9.2.1 IDEAL GAS EQUATION

The evaluation of the properties of an ideal gas will now be considered. It has been shown that the internal energy and enthalpy of an ideal gas are not functions of the volume or pressure, and hence these properties are simply functions of temperature alone. This means that the specific heats at constant volume and constant pressure are not partid derivatives of temperature, but can be written

Also, if c, and cp in molar quantities are denoted by cv,, and cp,,, then, for ideal and perfect gases

(9.8) It is possible to evaluate the properties of substances in terms of unit mass (specific

properties) or unit amount of substance (molar properties). The former will be denoted by lower case letters (e.g. v, u, h, g , etc), and the latter will be denoted by lower case letters and a suffix m (e.g. v,, u,, h,, g,, etc). The molar properties are more useful for combustion calculations, and these will be considered here.

Cp.m - c v , m = 3

Page 177: Advanced Thermodynamics for Engineers

State equation for ideal gases 161

The molar internal energy and molar enthalpy can be evaluated by integrating eqns (9.7), giving

(9.9)

where uO,, and h0,, are the values of u, and h, at the datum temperature To. Now, by definition

h, = U, + pv, and for an ideal gas pv, = %T; thus

h,=u,+%T (9.10) Hence, at T = 0, u, = h,, i.e. u0,, = ho,,,. (A similar relationship exists for the specific properties, u and h.)

To be able to evaluate the internal energy or enthalpy of an ideal gas using eqn (9.10) it is necessary to know the variation of specific heat with temperature. It is possible to derive such a function from empirical data by curve-fitting techniques, if such data is available. In some regions it is necessary to use quantum mechanics to evaluate the data, but this is beyond the scope of this course. It will be assumed here that the values of cv are known in the form

(9.1 1) where a , b, c and d have been evaluated from experimental data. Since cp,,- cV,,,=% then an expression for cp,, can be easily obtained. Hence it is possible to find the values of internal energy and enthalpy at any temperature if the values at To can be evaluated. This problem is not a major one if the composition of the gas remains the same, because the datum levels will cancel out. However, if the composition varies during the process it is necessary to know the individual datum values. They can be measured by calorimetric or spectrographic techniques. The ‘thermal’ part of the internal energy and enthalpy, i.e. that which is a function of temperature, will be denoted

cV., = a + bT + cT2 + dT3

(9.12)

(9.13)

assuming that the base temperature is absolute zero. Enthalpy or internal energy data are generally presented in tabular or graphical form.

The two commonest approaches are described below. First, it is useful to consider further the terms involved.

9.2.2 As previously discussed, uO,, and h0,, are the values of molar internal energy, u,, and molar enthalpy, h,, at the reference temperature, To. If To = 0 then, for an ideal gas,

THE SIGNIFICANCE OF uO,, AND h0.,

(9.14)

Page 178: Advanced Thermodynamics for Engineers

162 Thermodynamic properties of ideal gases and ideal gas mixtures

!f To # 0, then uO,, and ho,m are different. Most calculations involve changes in enthalpy or mtemal energy, and if the composition during a process is invariant the values of ho,, or u0 .m will cancel.

However, if the composition changes during a process it is necessary to know the difference between the values of u0,,, for the different species at the reference temperature. This is discussed below.

Obviously u0,, and h0,, are consequences of the ideal gas assumption and the equation

hm = hm(T) + h0,m contains the assumption that the ideal gas law applies down to To. If To=O, or a value outside the superheat region for the gas being considered, then the gas ceases to be ideal, often becoming either liquid or solid. To allow for this it is necessary to include latent heats. This will not be dealt with here, but the published data do include these corrections.

9.2.3 ENTROPY OF AN IDEAL GAS - THIRD LAW OF THERMODYNAMICS

The change in entropy during a process is defined as

If the functional relationship h, = h,(T) is known giving

(9.15)

then eqn (9.15) may be evaluated,

(9.16)

where so,,, is the value of s, at To and po.

shown that It is convenient to take the reference temperature To as absolute zero. It was previously

(9.17)

Consider

T-0 lim ($) If cp,,(T0) 3 0 then the expression is either infinite or indeterminate. However, before reaching absolute zero the substance will cease to be an ideal gas and will become a solid. It can be shown by the Debye theory and experiment that the specific heat of a solid is given by the law cp = aT3. Hence

lim (4") - = T ~ o lim ( - = pz (UT2 dT)+O T + O

(9.18)

To integrate from absolute zero to T it is necessary to include the latent heats, but still it is possible to evaluate jdh,/T.

Page 179: Advanced Thermodynamics for Engineers

If s,(T) is defined as

then entropy

State equation for ideal gases 163

(9.19)

(9.20)

The term so,, is the constant of integration and this can be compared with u0,,, and ha,. If the composition is invariant then the value of so,,, will cancel out when evaluating changes in s and it is not necessary to know its value. If the composition varies then it is necessary to know, at least, the difference between so values for the substances involved. It is not possible to obtain any information about so from classical (macroscopic) thermodynamics but statistical thermodynamics shows that 'for an isothermal process involving only phases in internal equilibrium the change in entropy approaches zero at absolute zero'. This means that for substances that exist in crystalline or liquid form at low temperatures it is possible to evaluate entropy changes by assuming that differences in so,, are zero.

It should also be noted that the pressure term, p , in eqn (9.20) is the partial pressure of the gas if it is contained in a mixture, and po is a datum pressure (often chosen as 1 bar or, in the past, 1 atmosphere).

9.2.4

The Gibbs energy will now be derived for use later when considering the equilibrium composition of mixtures (dissociation). By definition,

g,= h,- Ts, (9.21)

THE GIBBS ENERGY (FUNCTION) OF AN IDEAL GAS

and

h m = hm(T) + h0.m (9.22)

If the pressure ratio, PIPo, is denoted by the symbol p r , i.e. p r = p / p o , then

s, = s,(T) - 3 In p r + so,,, (9.23)

and

g = h(T) + h, - T(s(T) - 3 In p r + so)

= (h, - Ts,) + (h(T) - Ts(T)) + 3 T In p , (9.24)

If the terms at To are combined to give

g 0 . m = h ~ , m - Ts0.m (9.25)

and the temperature-dependent terms are combined to give

grn(T) = h m ( T ) - TSm(T) then

(9.26)

Page 180: Advanced Thermodynamics for Engineers

164 Thermodynamic properties of ideal gases and ideal gas mixtures

Often g,(T) and go,,, are combined to give g i which is the pressure-independent portion of the Gibbs energy, i.e. g i = g,(T) + go,,, and then

gm= g: + 9tT Pr (9.28)

g: is the value of the molar Gibbs energy at a temperature T and a pressure ofp , , and it is a function of temperature alone. The datum pressure, p o , is usually chosen as 1 bar nowadays (although much data are published based on a datum pressure of 1 a m ; the difference is not usually significant in engineering problems, but it is possible to convert some of the data, e.g. equilibrium constants).

Gibbs energy presents the same difficulty when dealing with mixtures of varying composition as u,, h, and so. If the composition is invariant, changes in Gibbs energy are easily calculated because the go terms cancel. For mixtures of varying composition go must be known. It can be seen from eqn (9.26) that if the reference temperature, To, equals zero then

g 0 , m = h0.m = U0.m (9.29)

9.3 Tables of u ( T ) and h ( T ) against T

These tables are based on polynomial equations defining the enthalpy of the gas. The number of terms can vary depending on the required accuracy and the temperature range to be covered. This section will limit the number of coefficients to six, based on Benson (1977). The equation used is of the form

(9.30)

The values of the coefficients for various gases are listed in Table 9.3, for the range 500 to 3000 K. If used outside these ranges the accuracy of the calculation will diminish.

Table 9.3 Enthalpy coefficients for selected gases found in combustion processes (based on W / kmol, with the temperature in K )

h, = a, Substance as a4 a3 a2 a1 a6 kJ/kmOl

O.oo00 O.oo00 O.oo00 O.oo00 O.oo00 O.oo00 O.oo00 -8.5861 le-15 O.oo00

- 1.44392e-11 -2.1945Oe-12 -6.5747Oe-12 -4.9036Oe-12

8.66002e-11 1.53897e-11

1.62497e-10 - 1.8180% 1 1

- 1.3867Oe- 1 1

9.6699Oe-8 -3.2208Oe-8

1.95300L-9 -9.58800~-9 -7.88542e-7

4.95240~-8 - 1.24402e-6

1 .@187e-7

- 1.49524e-7

-8.18100L-6 3.7697Oe-4 2.9426Oe-4 2.9938Oe-4 2.731 14e-3 6.5235Oe-4 5.65590S-4 4.96462e-3

-2.51427e-4

3.43328 - 3.317000 3.34435 3.50 174 3.09590 3.25304 3.74292 1.93529 2.76403

-3.84470 4.63284 3.75863 5.11346 6.58393 5.71243 9.6514Oe- 8.15300 3.73309

O.oo00

O.oo00 8.99147~~4

-3.93405e5 O.oo00

1 -2.39082e5 -6.6930~4

- 1.13882e5

2.46923e5

Hence the enthalpy, internal energy, enwopy and Gibbs energy can be evaluated as follows.

Page 181: Advanced Thermodynamics for Engineers

Tables of u( T ) and h( T ) against T 165

Enthalpy

h,(T) = %T(a , + a2T + a3T2 + a4T3 + a,T4) = %(a,T + a2T2 + a,T3 + a4T4 + a5TS)

c

= % 1 aiTi i = 1

Internal energy

u J T ) = % T ( ( a , - 1 ) + a2T + a,T2 + a4T3 + a5T4) = % ( ( a , - l )T + a2T2 + a,T3 + a4T4 + a,T5)

5

= % aiTi - %T = h(T) - %T

(9.31)

(9.32) i = 1

Entropy

This is defined by eqn (9.15) as

Integrating eqn (9.15) gives

Equation (9.20) provides some problems in solution. The first is that

and this results in ln(0) when T = To. Fortunately these problems can be overcome by use of the Van't Hoff equation, which will be derived in Chapter 12, where dissociation is introduced. This states that

where, for To = 0 K

p: = chemical potential = g,(T) + go,m = gm(T) + h0.m

(9.33)

(9.34)

and

hm = hm(T) + h0,m

(Note: the term &, is similar to g:, which was introduced in eqn (9.28). At this stage it is sufficient to note that the chemical potential has the same numerical value as the specific Gibbs energy. The chemical potential will be defined in Section 12.2.)

Page 182: Advanced Thermodynamics for Engineers

166 Thermodynamic properties of ideal gases and ideal gas mixtures

By definition, in eqn (9.31),

giving

-=- kdT) a' +a, + q T + q T 2 + 4 T 3 RT T

Substituting eqns (9.34), (9.13) and (9.36) into eqn (9.33) and rearranging gives

gm(T) + h0.m ) RT2

which can be integrated to

a3T2 a4T3 a5T4 gm(T) + h0.m In T + a,T+ - + - + - - - 2 3 4 RT 9lT

It is conventional to let

A = a, - a6

and then

3 4

The value of entropy can then be obtained from

giving

4 3 4 2 3

h T + h 2 T + - %T2 + - a,T3 +

Gibbs energy

The Gibbs energy can be evaluated from eqn (9.40) or eqn (9.26):

(9.35)

(9.36)

(9.37)

(9.38)

(9.39)

(9.40)

(9.41)

(9.42)

The values obtained from this approach, using eqns (9.31) to (9.42) are given in Tables 9.4 for oxygen, nitrogen, hydrogen, water, carbon monoxide, carbon dioxide, nitric oxide and methane. The tables have been evaluated up to 3500 k, slightly beyond the range stated in relation to Table 9.3, but the error is not significant. This enables most combustion problems to be solved using this data. Other commonly used tables are those of JANAF (1971).

Page 183: Advanced Thermodynamics for Engineers

Table 9.4 Molecular oxygen, 0, : ho = O.oo00 kJ/kmol

T Yr, u(r, Str, fir, 50 1365.74 950.03 153.84 4326.3 100 2757.68 1926.25 173.12 -14553.9

200 5616.56 3953.70 192.89 -32962.1 250 7081.75 5003.18 199.43 -42775.9 298 85M.70 6032.04 201.65 -52477.3 300 8569.64 6075.35 204.85 -52886.8 350 10079.40 7169.40 209.51 -63248.6 400 11610.23 8284.51 213.60 -73828.2

150 4174.92 2927.77 184.60 -23515.4

450 500 550 600 650 700 750 800 850 900 950

lo00

-

-

13161.33 1473 1.93 16321.30 17928.68 19553.39 21 194.71 2285 1.99 24524.55 2621 1.78 27913.05

94 19.89 10574.78 11748.43 12940.10 14149.09 15374.70 16616.26 17873. I 1 19144.62 20430.18

217.25 220.56 223.59 226.39 228.99 231.42 233.71 235.86 237.91 239.85

-84601.0 -95547.5 -106652.2 -1 17902.4 -129287.4 -140798.2 - 152426.8 - I64 166.6 -17601 1.4 -1 87955.8

29627.76 21729.18 241.71 -199995.3 31355.35 23041.05 243.48 -212125.3

I050 33095.25 24365.23 245.18 -224342.1 1100 34846.92 25701.19 246.81 -236642.0 I150 36609.84 27048.40 248.38 -249021.9 I200 38383.52 28406.36 249.89 -261478.6 1250 40167.47 29774.59 251.34 -274009.5 1300 41961.22 31152.63 252.75 -286612.0 1350 43764.34 32540.04 254.11 -299283.6 1400 45576.41 33936.39 255.43 -312022.2 1450 47397.01 35341.28 256.71 -324825.7 I500 49225.76 36754.31 257.95 -337692.1 1550 51062.30 38175.14 259.15 -350619.6 1600 52906.28 39603.40 260.32 -363606.5 1650 54757.37 41038.78 261.46 -376651.1 1700 56615.26 42480.95 262.57 -389752.0 1750 58479.67 43929.65 263.65 402907.6 1800 60350.32 45384.58 264.70 -416116.5 1850 62226.96 46845.50 265.73 -429377.5 1900 64109.36 48312.19 266.74 -442689.3 1950 65997.31 49784.42 267.72 -456050.7 2000 67890.61 51262.01 268.68 -469460.6 2050 69789.09 52744.78 269.61 -482918.0 2100 71692.60 54232.57 270.53 -494421.6 2150 73600.99 55725.25 271.43 -509970.7 2200 75514.16 57222.70 272.31 -523564.2 2250 77432.00 58724.83 273.17 -537201.2

2350 81281.41 61742.81 274.84 -564602.5 2400 83212.89 63258.57 275.66 -578365.1 2450 85148.84 64778.80 276.46 -592168.0 2500 87089.26 66303.51 277.24 -606010.4 2550 89034.18 67832.71 278.01 -619891.8 2600 90983.62 69366.44 278.77 -633811.2 2650 92937.65 70904.76 279.51 447768.3 2700 94896.35 72447.74 280.24 -661762.2 2750 96859.80 73995.47 280.96 -675792.5 2800 98828.1 I 75548.07 281.67 489858.5 2850 100801.43 77105.68 282.37 -703959.7 2900 102779.90 78668.43 283.06 -718095.5 2950 104763.70 80236.52 283.74 -732265.5 3000 106753.02 81810.12 284.41 -746469.2 3050 108748.05 83389.44 285.07 -760706.1 3100 110749.05 84974.72 285.72 -774975.8 3150 112756.24 86566.20 286.36 -789277.8 3200 114769.90 88164.14 286.99 -80361 1.7 3250 116790.32 89768.85 287.62 -817977.1 3300 118817.80 91380.61 288.24 -832373.6 3350 120852.67 92999.76 288.85 -846800.9 3400 122895.26 94626.64 289.46 -861258.7 3450 124945.95 96261.62 290.06 -875746.5 3500 127005.11 97905.06 290.65 -890264.2

2300 79354.44 60231.55 274.02 -550880.9

Molecular nitrogen, N, : ho = O.oo00 kJhol T 40 u(n Stn dn 50 1396.41 980.70 140.27 4617.2 100 2805.07 1973.64 159.79 -13174.0 150 4225.96 2978.82 171.31 -21470.5 200 5659.09 3996.23 179.55 -30251.7 250 7104.43 5025.86 186.00 -393%.4 298 8503.43 6025.77 191.12 -48450.8 300 8561.97 6067.68 191.32 -48833.3 350 10031.66 7121.65 195.85 -58515.2 400 11513.46 8187.74 199.81 48408.6 450 13007.34 9265.90 203.32 500 14513.22 10356.07 206.50 550 16031.05 11458.18 209.39 600 17560.74 12572.16 212.05 650 19102.23 13697.93 214.52 700 20655.41 14835.40 216.82 750 22220.19 15984.47 218.98 800 23796.47 17145.03 221.01 850 25384.12 18316.97 222.94 900 26983.03 19500.16 224.77 950 28593.06 20694.47 226.51

loo0 30214.07 21899.77 228.17

450 500 550 600 650 700 750 800 850 900 950

loo0

-

-

13007.34 14513.22 1603 1.05 17560.74 19102.23 20655.41 22220.19 23796.47 25384.12 26983.03 28593.06 30214.07

9265.90 10356.07 1 1458.18 12572.16 13697.93 14835.40 15984.47 17145.03 183 16.97 19500.16 20694.47 2 1899.77

203.32 206.50 209.39 212.05 214.52 216.82 218.98 221.01 222.94 224.77 226.51 228. I7

1050 I 1 0 0 1150 1200 1250 I300 1350 I400 1450 1500 I550 1600 1650 1 700 1750 1800 1850 1900 1950 2000 2050 2100 2150 2200 2250 2300 2350 2400

31845.92 33488.45 35141.49 36804.88 38478.43 40161.95 41855.26 43558.13 45270.36 46991.73 48722.01 50460.96 52208.33 53963.87 55727.32 57498.40 59276.85 61062.36 62854.64 64653.40 66458.32 68269.07 70085.34 71906.79 73733.06 75563.81 77398.68 79237.30

23115.90 24342.72 25580.04 26827.72 28085.55 29353.36 30630.95 31918.11 33214.63 34520.28 35834.85 37158.08 38489.74 39829.56 41177.30 42532.66 43895.39 45265.19 46641.76 48024.80 49414.00 50809.04 52209.60 53615.33 55025.89 56610.92 57860.08 59282.98

229.76 231.29 232.76 234.18 235.54 236.86 238.14 239.38

-78488.4 -88735.2 -99133.4 -109670.4 -120335.4 -131 119.6 -142015.2 -153015.5 -164114.8 -1 75307.9 -186590.2 -197957.5 -209406. I -220932.8 -232534.4 -244208.0 -255951.3 -267761.7 -279637.0 -291 575.2

-

240.58 -303574.5 241.75 -315632.9 242.88 -32n48.9 243.99 -339920.9 245.06 -352147.3 246. I I -364426.8 247.13 -376758.1 248.13 -389139.9 249. I I 401570.9 250.06 dl4050.2 -. . ..

250.99 -426576.5 251.90 -439148.9 252.79 253.67 254.52 255.36 256.18 256.98 257.77 258.55

-451766.3 -464427.8 -477132.5 -489879.5 402668.0 -5 15497.1 -528366.0 -541274.1

2450 81079.29 60709.25 259.31 -554220.4 2500 82924.26 62138.51 260.05 -567204.4 2550 84771.83 63570.36 260.78 -580225.3 2600 86621 58 6500440 261 50 -5932825 2650 88473 12 6644022 26221 -6063753 2700 9032601 6787740 26290 -6195030 2750 9217984 6931552 26358 -6326650 2800 94034 17 70754 13 26425 -6458608 2850 9588856 7219281 26490 -6590897 -..

2900 2950 3000 3050 3100 3150 3200 3250 3300 3350 3400 3450 3500

-

-

97742 56 73631 09 265 55 9959570 75068.51 266 18 10144752 7650462 26681 103297.54 105 145.28 106990.25 108831.95 110669.87 112503.49 114332.28 116155.72 117973.26 119784.36

77938.92 79370.95 80800.21 82226.19 83648.39 85066.30 86479.38 87887.10 89288.93 90684.3 I

267.42 268.02 268.61 269.19 269.76 270.32 270.87 271.41 271.94 272.46

-67235 I . I 685644.4 -698969.2 -712324.8 -725710.7 -739126.4 -752571.4 -766045. I -779547. I -793076.8 -806633.8 -820217.5 -833827.5

Page 184: Advanced Thermodynamics for Engineers

Table 9.4 continued Carbon monoxide, CO :ho = -1 13 882.3 k J h 0 l

T Mn NT, s(n B(0 50 138673 971 01 14672 -59492

100 278893 195750 166 I5 -138256 I50 420639 295924 17764 -224390 200 563890 397604 I85 88 -315363 250 708627 500769 19233 -409973 298 848951 601184 19747 -503560 300 854826 605397 19766 -50751 I 350 1002468 711467 20222 607508 400 1151529 818957 20620 -709632 450 1301987 927844 20974 -81363 2 500 1453821 1038106 21294 -919315 550 1607008 1149721 21586 -1026525 600 1761524 1262666 21855 -1135136 650 1917347 1376917 221 04 -124504 I 700 2074452 1492451 22337 -1356150 750 800 850 900 950

1000 1050 1100 I I50 1200 1250 1300 1350 I400

22328.17 23924. I6 25532.26 27152.23 28783.80 30426.74 32080.79 33745.68 35421.17 37106.98 38802.85 40508.52 42223.72 43948. I6

16092.44 17272.72 18465.1 1 19669.36 20885.22 22112.44 23350.77 24599.95 25859.72 27129.82 28409.98 29699.93 30999.41 32308.14

225.56 227.62 229.57 23 1.42 233.18 234.87 236.48 238.03 239.52 240.95 242.34 243.68 244.97 246.23

-146838.7 - 158 168.5 -169598.5 -181123.5 -192738.8 -204440.3 -2 16224.3 -228087.3 -240026.3 -252038.4 -264120.9 -276271.5 -288487.9 -300768.0

1450 45681.58 33625.84 247.44 -313109.8- 1500 47423.69 34952.24 248.62 -325511.6 1550 49174 22 36287 06 249 77 -337971 6 1600 50932 88 3763000 25089 -350488 3 1650 5269938 3898079 251 98 -3630600 1700 5447344 40339 13 25303 -3756853 1750 I800 1850 1900 I950 2000 2050 2100 2150 2200 2250 2300 2350 2400 2450 2500 2550 2600 2650 2700 2750 2800 2850 2900 2950 3000 3050 3100 3150 3200 3250 3300

56254 75 58043 03 59837 98 61639 29 63446 67 65259 81 67078 39 68902 I I 70730 65 72563 69 74400 92 76242 01 78086 64 79934 47 81785 19 83638 45 85493 91 87351 25 89210 I I 91070 16 68621 55 26993 92931 04 7006671 27062 94792 40 71512 36 271 29 9665390 72958 I5 27195 98515 I8 74403 71 272 59

100375 87 102235 62 10409407 7873545 27447 10595084 8017651 27507 10780557 8161553 27567 109657 89 83052 I3 276 25 I I I507 42 84485 95 276 82 11335379 8591660 27739

41704 73 43077 29 44456 53 45842 12 47233 79 48631 21 50034 07 51442 08 52854 90 54272 23 55693 74 571 19 12 58548 03 59980 I5

254 07 255.07 256.06 257.02 257.96 258.88 259.77 260.65 261.51 262.36 263.18 263.99 264.79 265.56

61415.15 266.33 62852.70 267.08 64292.45 267.81 65734.07 268.53 67177.22 269.24

~~

7584868 273 23 77292 72 273 86

-388363.0 -401091.6 -413870.1 -426697.1 -439571.6 -452492.6 -465458.9 -478469.7 -491523.9 -504620.8 -5 17759.3 -530938.7 -5441 58.2 -557417.0 -57071 4.3 -584049.4 -597421 5 -610830. I 424274.4 437753.9 451267.7 -664815.4 478396.4 -6920 10.0 -705655.7 -719332.9 -733041. I -746779.8 -760548.4 -774346.4 -788173 3 402028.7

3350 1151%61 8734371 27794 -815912 I 3400 I17035 50 88766 88 278 49 -829822 9 3450 I1887008 90185 75 27902 -8437607 3500 12069996 9159991 27955 -8577250

Carbon dioxide, C02 : ho = -393 404.9 M h o J T m 4n m dn

50 1342.97 927.25 157.68 -6541.2 100 2794.62 1%3.19 177.72 -14977.7 150 4350.19 3103.05 190.31 -24196.3 200 6005.05 4342.19 199.82 -33958.6 250 7754.65 5676.08 207.62 -44150.0 298 9519.29 7041.63 214.07 -54273.9 300 9594.56 7100.27 214.32 -54702.3 350 400 450 500 550 600 650 700 750 800 850 900 950

lo00 -.~. ~~ . 1050 45347.89 36617.87 1100 48118.27 38972.54 1150 50920.17 41358.72 I200 53751.08 43773.92 1250 56608.63 46215.76 1300 59490.55 48681.96 1350 62394.68 51170.37 1400 65318.96 53671194 1450 1500 1550 74193.77 1600 77180.26 1650 80178.25 I700 83186.33 1750 86203.20 I800 89227.66 1850 92258.64 1900 95295.17 I950 98336.36 2000 101381.47 2050 104429.85 2100 107480.95 2150 110534.33 2200 113589.67 2250 116646.75 2300 119705.46 2350 122765.79 2400 125827.86 2450 128891.87 108521.83 321.97 2500 131958.14 111172.39 323.21 2550 135027.11 113825.64 324.42 2600 138099.31 116482.13 325.62 2650 141175.38 119142.48 326.79 2700 144256.08 121807.47 327.94 2750 147342.28 124477.95 329.07 2800 150434.93 127154.89 330.19 2850 153535.12 129839.36 331.28 2900 156644.02 132532.55 332.37 2950 159762.95 135235.76 333.43 3000 162893.29 137950.39 334.48 3050 166036.55 140677.93 335.52 3100 169194.35 143420.02 336.55 3150 172368.42 146178.38 337.57 3200 175560.59 148954.83 338.57 3250 178772.80 151751.32 339.57 3300 182007.10 154569.91 340.55

11520.46 13528.14 15613.47 17772.48 20001.26 22296.03 24653.12 27068.% 29540.09 32063.16 34634.93 37252.26

8610.46 10202.42 11872.04 13615.33 15428.39 17307.45 19248.83 2 1248.95 23304.37 2541 1.72 27567.78 29769.39

220.26 225.62 230.53 235.08 239.32 243.32 247.09 250.67 254.08 257.33 260.45 263.44

39912.12 32013.54 266.32 4261 1.60 34297 30 269.09

271.76 274.34 276.83 279.24 281.57 283.83 286.02 288.15 290.21 292.22

61306.60 294.17 63877.38 2%.07 66159.66 297.91 69052.02 299.71 71653.17 301.46 74261.92 303.16 76877.19 304.82 79498.00 306.44 82123.48 308.02 84752.87 309.56 87385.53 311.07 90020.92 312.54 92658.58 313.98 95298.21 315.38 97939.57 316.75

100582.57 318.10 103227.19 319.42 105873.54 320.70

68261 44 56205 70 71220 29 58748 84

45569.5 -767 18.5 -88123.7 -99765.1

-1 11626.2 -123693.2 -135954.1 -148398.8 -16lO18.1 -173804.0 -186749.2 -199847.1 -213091.6 -226477.3 -239998.8 -25365 I .6 -267431.0 -281332.9 -295353.4 -309488.7

-338089.8 -352549.2

-323735.3

-367110.3

-396526.4 -4 11376. I -4263 16.8 -441346.1 -456461.7 -471661.4 -486943.2 -502304.9 -517744.7

-548850.9

-381770.3

-533260.6

-564513.9 -580248.0 -596051.5 -611923.0 -627861.0 -643864. I -659931.0 -676060.5 492251.4 -708502.5 -724812.6 -741180.9 -757606.3 -774087.8 -790624 7 -8072 16.0 -823861.0 -840558.9 -857309.2 -874 I 1 1 .O -890964.0 -907867.5 -924821.0 -94 1824.0 ~~ ~ . .

3350 185265.64 157412.74 341.53 -958876.3 3400 188550.70 160282.08 342.51 -975977.4 3450 19186464 163180.30 343.48 -993127.0 3500 195209.95 166109.90 344.44 -1010324.9

Page 185: Advanced Thermodynamics for Engineers

Table 9.4 continued

Molecular hydrogen, H2 : ho = 0.0000 k l h o l T Mn d7) s(n

50 1427.20 1011.48 79.70 -2557.8 100 285464 2023 21 9949 -70942 150 428292 303578 11107 -123776 200 571258 404972 11930 -181466 250 7144 17 506560 12568 -242770 298 8520.79 6043.13 130.72 -30434.3 300 8578.21 6083.92 130.91 -30695.9 350 1001520 7105 19 13534 -37355 1 400 1145563 812991 I39 I9 -442206 450 12899% 915853 14259 500 1434865 10191 50 I4565 550 15802 13 1122926 14842 600 1726081 1227223 15095 650 1872508 1332079 15330 700 2019534 1437533 15548 750 2167192 1543620 15751 800 23155 19 16503 75 15943 850 2464546 1757830 161 24 900 2614303 18660 16 16295

-51266.8 -58474.1 -65826.7 -733 11.8 -80918.9 -88638.9 -96464.3

-104388.4 -112405.4 -1 205 10.4

950 27648.20 19749.61 164.58 -128698.8 1000 29161.23 20846.93 166.13 -136966.7 1050 1100 1150 I200 1250 1300 1350 1400 1450 1500 1550 I600 1650 1700 1750 1 800 1850 1900 1950 2000 2050 2100 2150 2200 2250 2300 2350 2400 2450 2500 2550 2600 2650 2700 2750 2800 2850 2900

30682.38 32211.88 33749.95 352%.78 36852.56 38417.44 39991.57 41575.07 43168.06 44770.62 46382.83 48004.73 49636.37 51277.76 52928.90 54589.77 56260.35 57940.56 59630.34 61329.61 63038.25 64756.13 66483. I 1 68219.03 69963 71 71716.95 73478.53 75248.23 77025.78 78810.91 80603.35 82402.77 84208.86 86021.28 87839.67 89663 63 91492.79 93326.72

21952.37 23066. I5 24 188.5 1 253 19.62 26459.68 27608.85 28767.26 29935.05 311 12.33 32299.17 33495.66 34701.85 35917.77 37143.45 38378.88 39624.03 40878.89 42143.39 43417.46 44701.01

167.61 169.04 170.40 171.72 172.99 174.22 175.41 176.56 177.67 178.76 179.82 180.85 181.85 182.83 183.79 184.73 185.64 186.54 187.42 188.28

45993.93 472%. IO 48607.36 49927.57 51256.54 52594.06 53939.93 55293.91 56655.74 58025.16 59401.88 60785.59 62175.97 63572.67 64975.34 66383.59 67797.04 69215.25

189.12 189.95 190.76 191.56 192.34 193.11 193.87 194.62 195.35 196.07 196.78 197.48 198.17 198.84 199.51 200.17 200.82 201.45

-145310.5 -1 53726.9 - 1622 13. I -170766.4 -179384.3 -188064.6 -1%805.3 -205604.5 -2144604 -223371 5 -232336 I -241352 9 -250420 5 -259537 7 -268703 4 -277916 4 -287175 6 -296480 1 -305829 0 -315221 4 -324656 3 -334133 0 -343650 8 -353208 8 -362806 3 -372442 8 -3821 I7 4 -391829 6 -401578 8 -41 I364 3 -421 I85 6 -431042 I -440933 3 -450858 6 -4608176 -470809 6 480834 3 -490891 I

2950 9516500 7063782 20208 -5009796 3000 9700717 7206427 20270 -5110993 3050 98852 76 73494 I5 203 31 -521249 7 3100 100701 29 74926% 20391 -5314303 3150 10255225 7636221 20451 -5416409 3200 104405 12 7779936 20509 -5518808 3250 106259 36 79237 89 205 66 -562149 6 3300 10811442 8067723 20623 -572447 I 3350 10996970 8211680 20679 -5827726 3400 I I1824 63 83556 01 207 34 -593125 8 3450 11367859 8499425 20788 -6035063 3500 11553094 8643089 20841 -6139136

Water, H 2 0 : ho = -239 081.7 kJhol T Mn u(n s(r) m

50 1567.79 1152.08 130.24 -4944.1 100 315940 232797 15228 -120689 150 477508 352794 16538 -20031 6 200 6415 I O 475224 17481 -28547 I 250 807969 6001 I I 18224 -374797 298 97W99 722333 18817 -463733 300 976904 727475 18840 -467499 350 11483 36 8573 35 19368 -563049 400 1322278 989706 19833 -661073 450 1498747 1124603 20248 -761293 500 1677752 1262037 20625 -86349 I 550 1859304 14020 18 20971 -%7495 600 20434 IO 15445.52 212.92 -107316.2 650 22300.74 168%.45 215.91 -118037.6 700 24192.99 18372.98 218.71 -128903.7 750 26110.85 19875.13 221.36 -139905.9 800 28054.30 21402.86 223.86 -151036.9 850 30023.31 22956.15 226.25 -162290.3 900 32017.79 24534.92 228.53 -173660.3 950 34037.68 26139.09 230.72 -185141.8

1000 36082.85 27768.55 232.81 -1%730.4 IO50 38153.17 29423.15 234.83 -208421.8 1 100 40248.48 31102.75 236.78 -220212.5 1150 42368.62 32807 17 238.67 -232099.0 I200 44513.37 34536.21 240.49 -244078.3 I250 46682.51 36289.64 242.26 -256141.4 1300 48875.80 38067.21 243.98 -268303.8 1350 51092.97 39868.66 245.66 -280545.1 1400 53333.72 41693.70 247.29 -292868.9 1450 55597.74 43542.00 248.88 -305273.2 1500 57884.69 45413.24 250.43 -3in55.9 1550 60194.21 47307.04 251.94 -330315.3 1600 62525.92 49223.04 253.42 -342949.5 1650 64879.42 51 160.82 254.87 -355656.9 1700 67254.27 53119.96 256.29 -368436.0 1750 6%50.03 55100.00 257.68 -381285.3 1800 72066.22 57100.48 259.04 -394203.3 1850 74502.34 59120.89 260.37 -407188.7 1900 76957.88 61160.71 261.68 -420240.2 1950 79432.30 63219.41 262.97 -433356.6 2000 81925.03 652%.43 264.23 -446536.7 2050 84435.48 67391.17 265.47 -459779.4 2100 86%3.05 69503.02 266.69 -473083.4 2150 89507.10 71631.36 267.89 -486447.9 2200 92066.98 7377552 269.06 -499871.7 2250 94642.01 75934.83 270.22 -513353.9 2300 97231.49 78108.60 271.36 -526893.4 2350 99834.70 802%.09 272.48 -540489.4 2400 102450.89 824%.57 273.58 -554141.0 2450 105079.29 84709.26 214.66 -567847.1 2500 107719.12 86933.37 275.73 -581607.1 2550 2600 2650 2700 2750 2800 2850 2900 2950 3000 3050 3100 3150 3200 3250 3300 3350 3400 3450 3500

110369.56 113029.77 115698.90 118376.06 121060.35 123’750.84 126446.58 129146.60 131849.91 134555.49 137262.29 139969.26 I42675 3 I 145379.33 148080.18 150776.73 153467.78 156152.15 158828.60 161495.91

89168. IO 91412.59 93666.00 95927.45 981%.02

100470.80 102750.82 105035.13 107322.73 109612.59 I 1 1903.68 114194.93 116485.26 118773.57 121058.71 123339.54 125614.88 127883.53 130144.27 132395.86

276 78 277 81 278 83 279 83 280 82 281 79 282 74 283 68 284 60 285 51 28641 287 29 288 15 289 01 289 84 290 67 291 48 292 27 293 05 293 82

-595419.9 -609284.8 -623200.9 -637167.5 -651183.8 -665248.9 -679362. I -693522.6 -707729.7 -72 1982.7 -736280.8 -750623.2 -765009.3 -779438.4 -793909.6 -808422.4 -822976.0 -837569.8 -852202.9 -866874.8

Page 186: Advanced Thermodynamics for Engineers

Table 9.4 continued Nitric oxide, NO : ho = 89 9 14.7 klkmol T YT) u(T) Jin dT)

50 1461.94 1046.22 156.66 -6371.1 100 293626 210483 17709 -147726 150 4422 89 317575 189 14 -239482 200 5921 77 425891 19776 -336309 250 743280 535422 20451 436936 .. . . 298 8894.74 6417.08 209.85 -53641.9 300 8955.89 6461.60 210.06 -54061.8 350 10490.97 7580.97 214.79 61685.9 400 12037.92 8712.20 218.92 -75530.9 450 13596.65 9855.21 222.59 -86570.5 500 15167.03 11009.88 225.90 -97784.3 550 600 650 700 750 800 850 900

16748.95 18342.30 19946.93 2 1562.7 I 23189.50 24827.15 26475.50 281 34.40

121 76.09 13353.72 14542.63 15742.70 16953.77 18175.71 19408.35 20651.53

228.92 231.69 234.26 236.65 238.90 241.01 243.01 244.91

-109155.9 -1 20672.0 -132321.5 -144095.0 -155984.4 -167982.7 -180083.8 -192282.1

950 29803.68 21905.10 246.71 -204573.0 1000 31483.16 23168.86 248.44 -216952.0 I050 33172.68 24442.66 250.08 -229415.2 1100 34872.03 25726.30 251.66 -241959.2 I I50 36581.03 27019.58 253.18 -254580.7 I200 38299.48 28322.32 254.65 -267276.7 1250 40027.19 29634.31 256.06 -280044.5 1300 41763.93 30955.34 257.42 -292881.6 1350 43509.49 32285.19 258.74 -305785.7 1400 45263.66 33623.64 260.01 -318754.6 1450 47026.20 34970.47 261.25 -331786.4 1500 48796.88 36325.43 262.45 -344879.0 I550 1600 1650 1 700 1750 1800 1850 1900

50575.47 52361.70 54155.33 55956.1 1 57763.77 59578.03 61398.62 63225.26

37688.30 39058.82 40436.74 4 I821 .SO 432 13.74 44612.29 46017.17 47428.09

263.62 264.75 265.86 266.93 267.98 269.00 270.00 270.97

-358030.9 -37 1240.2 -384505.5 -397825.2 41 1198.0 -424622.6 438097.7 451622.0

1950 65057.66 48844.77 271.92 465194.5 2000 66895.51 50266.91 272.85 -478814.1 2050 68738.53 51694.22 273.76 -492479.7 2100 70586.41 53126.38 274.66 -506190.3 2150 72438.82 54563.08 275.53 -519944.9 2200 74295.46 56004.00 276.38 -533742.7 2250 76155.99 57448.81 277.22 -547582.7 2300 78020.08 58897.19 278.04 -561464.1 2350 79887.40 60348.79 278.84 -575386. I 2400 81757.59 61803.27 279.63 -589347.8 2450 83630.32 63260.29 280.40 -603348.6 2500 85505.22 64719.47 281.16 -617387.6 2550 2600 2650 2700 2750 2800 2850 2900

87381.94 89260.09 91139.32 9301 9.23 94899.44 %779.56 98659.18

100537.91

66180.47 67642.91 69106.42 70570.62 72035.1 I 73499.52 74%3.43 76426.44

281.90 282.63 283.35 284.05 284.74 285.42 286.08 286.73

-631464.0 -645577.4 -659726.8 -673911.7 -68813 I .4 -702385.4 -716672.8 -730993.3

2950 102415.33 77888.15 287.38 -745346.1 3000 104291.02 79348.12 288.01 -759730.8 3050 3100 3150 3200 3250 3300 3350 3400 3450 3500

106164.57 108035.53 109903.48 1 11767.97 113628.55 115484.78 117336.18 119182.30 121022.66 122856.78

80805.95 82261.20 83713.43 85162.21 86607.08 88047.59 89483.28 90913.68 92338.32 93756.73

288.63 289.24 289.83 290.42 291.00 291.56 292.12 292.67 293.21 293.73

-774146.7 -788593.3 -803070.0 -817576.4 -8321 11.8 -846675.9 -861268.1 -875887.8 -890534.7 -905208.2

Methane, CH4 : ho = -66 930.5 kJho l T YT) u(T) JiT) n(T)

50 906.44 490.72 134.82 -5834.7 100 2011.62 1180.19 149.99 -12987.2 150 3308.10 2060.95 160.45 -20759.5 200 4788.60 3125.74 168.94 -29000.2

298 8197.36 5719.70 182.73 -46255.4 250 6446.08 4367.50 176.33 -37635.7

300 8273.64 5779.35 182.98 -46621.1 350 10264.62 7354.62 189.11 -55925.6 400 12412.51 9086.m 194.85 -65526.5 450 500 550 600 650 700 750 800 850 900 950

1000 I050 I 100 1150 1200 1250 1300 1350 1400

14711.00 17153.96 19735.44 22449.65 25291.00 28254.05 31333.53 34524.35 3782 1.57 41220.40 44716.23 48304.59 51981.17 55741.80 59582.46 63499.28 67488.52 71546.59 75670.02 79855.49

10969.57 200.26 -75406.0 12996.81 205.41 -85549.6 15162.57 210.33 -95945.2 17461.07 215.05 -106582.5 19886.70 219.61 -1 17452.3 22434.04 224.00 -128546.8 25097.81 ~ 8 . ~ 6 -139858.8 27872.91 232.38 -151381.7 30754.41 236.39 -163109.8 33737.53 240.29 -175037.6 36817.65 244.08 -187160.3 39990.29 247.78 -199473.3 4325 1.16 46596.07 50021.02 53522.12 57095.65 60738.00 64445.71 68215.47

251.38 254.91 258.34 261.71 265.00 268.22 271.37 274.46

-211972.4 -224653.9 -237514.1 -250549.7 -263757.9 -277135.7 -290680.7 -304390.6

1450 84099.80 72044.07 277.49 -318263.4 I500 88399.88 75928.43 280.46 -332297.0 1550 1600 1650 1700 1750 I800 1850 1900

92752.80 97 155.72

101605.94 106100.89 110638.09 115215.17 119829.90 124480.12

79865.63 83852.84 87887.35 91966.58 96088.06

100249.43 104448.44 108682.95

283.38 286.25 289.06 291.83 294.55 297.23 299.87 302.47

-346490.0 -360840.9 -375348.3 -39001 I .4 -404829.3 -419801.3 -434927.0 -450206.2

1950 129163.80 112950.91 305.03 -465638.8 2000 133878.99 117250.39 307.55 -481225.1 2050 138623.86 121579.55 310.04 496965.3 2100 143396.66 125936.63 312.50 -512860.1 2150 148195.74 130320.00 314.93 -528910.1 2200 153019.54 134728.08 317.33 -5451 16.4 2250 157866.57 139159.40 319.71 -561480.1 2300 162735.46 143612.57 322.06 -578002.6 2350 167624.89 148086.28 324.39 -594685.5 2400 172533.63 152579.31 326.69 -611530.6 2450 177460.54 157090.50 328.98 -628539.8 2500 182404.53 161618.78 331.25 -645715.4 2550 2600 2650 2700 2750 2800 2850 2900

187364.61 192339.82 197329.31 202332.27 207347.95 2 12375.65 217414.75 222464.67

166163.14 170722.64 1752%.42 179883.66 184483.62 189095.61 193719.00 198353.20

333.50 335.74 337.% 340.17 342.37 344.57 346.75 348.93

-663059.7 -680575.5 -698265.5

-734180.8 -752412.8 -770832.7 -789444.4

-716132.8

2950 227524.88 202997.69 351.11 -808252.1 3000 232594.89 207651.99 353.29 -827260.i 3050 3100 3150 3200 3250 3300 3350 3400 3450 3500

237674.27 242762.61

252964.82 258078.08 263199.09 268327.62 273456.95 278606.50 283756.50

247859.57

212315.65 216988.28 22 1669.52 226359.06 231056.60 235761.90 240474.72 245188.33 249922.16 254656.45

355.46 357.63 359.81 361.99 364.17 366.36 368.56 370.77 372.99 375.22

446473.L -865896.: -885534.f -905393.4 -925478. I -945794.7 -966349.4 -987145. I

-1008198.3 -1029506. I

Page 187: Advanced Thermodynamics for Engineers

Tables of u(T) and h(T) against T 171

9.3.1

Sometimes data are given in terms of mean specific heat rather than the actual specific heat at a particular temperature. This approach will now be described.

TABLES OF MEAN SPECIFIC HEAT

Consider the change of enthalpy, h, between temperatures T, and T2. This is

(9.43) Tz T

h2 - hl = J c,(T) dT = h(T) - h(TJ

Now, this can be written

C,(TZ - TI) = W 2 ) - W,) (9.44) where E, is the mean specific heat between T, and T2. Normally E , is given at a particular temperature T , and it is then defined as the mean between the temperature T and a reference temperature TEf, Le.

(9.45) T - T'L

It is also possible to write the mean specific heat as a polynomial function of temperature, in which case

(9.46)

If it is required to calculate the enthalpy difference between two temperatures T, and T2,

( E , ) = a + bT + cT2 + ... where a, b and c are tabulated coefficients.

then

h2 - h l = (Ep)T2(T2 - T ~ f ) - ( E p ) T ~ ( T I - T m f ) (9.47)

Having calculated the change in enthalpy, h, - h,, then the change in internal energy, u2 - u,, may be evaluated as

(9.48a) ~2 - u,= h2 - RT2 - (h1- RT,) = h, - h1- R(T2 - T , )

Page 188: Advanced Thermodynamics for Engineers

172 Thermodynamic properties of ideal gases and ideal gas mixtures

or, in molar terms, by

um.2 - u m , l = hm,2 - %Tz - (hm,] - % T I ) = hm,~ - hm.1- %(TZ - TI) (9.48b)

If the change of internal energy, u2 - u l , is known, then the change of enthalpy, h, - h, , may be calculated in a similar way. The use of mean specific heat values enables more accurate evaluation of temperature changes because the mean specific heats enable the energy equation to be solved with an allowance for the variation of specific heats. The mean specific heats approach is depicted in Fig 9.1.

9.4 Mixtures of ideal gases Many problems encountered in engineering involve mixtures of gases - air itself is a mixture of many gases although it can be considered to be oxygen and atmospheric nitrogen (which can be assumed to contain the nitrogen in the air and the other ‘inert’ gases such as argon and carbon dioxide). If the gases in a mixture are at a temperature well above their critical temperature and a pressure below the critical pressure, they act as ideal gases. Expressions relating the properties of mixtures of ideal gases will now be formulated. These expressions are a direct consequence of the Gibbs-Dalton laws.

9.4.1 DALTON PRINCIPLE

Dalton stated that

any gas is as a vacuum to any gas mixed with it.

This statement was expanded and clarified by Gibbs to give the Gibbs-Dalton laws.

9.4.2 GIBBS-DALTON LAW

[a) A gas mixture as a whole obeys the equation of state (eqn (9.2))

pV = n%T

where n is the total amount of substance in the mixture. (b) The total pressure exerted by a mixture is the sum of the pressures exerted by the

individual components as each occupies the whole volume of the mixture at the same temperature. The internal energy, enthalpy and entropy of the mixture are, respectively, equal to the sums of the internal energy, enthalpy and entropy of the various components as each occupies the whole volume at the temperature of the mixture.

[c)

9.4.3 MIXTURE RELATIONSHIPS

The total mass, m, of a mixture can be related to the mass of constituents by

(9.49)

From the ideal gas law

Page 189: Advanced Thermodynamics for Engineers

Mixtures of ideal gases 173

and hence for the individual constituents

pVa = maRaT = na%T pvb = m$bT = n,%T pV, = m P c T = n,%T, etc

(9.50)

In eqn (9.50), V , = volume of constituent a at the pressure and temperature of the mixture, and V , and V, are similar volumes for constituents b and c. But the total volume, V , is given by

n

v = v,+ v,+ v,+ ... =XI/; i = 1

and therefore n

n = na + n, + n, + = n, i = 1

Let

- xa = mole fraction of a in the mixture na _ - n

Similarly, nb/n =xb, and n,/n = x,, etc, with ni/n = xi. Therefore n

x, f xb + x, + . ’. = 4 = 1 i = 1

(9.51)

(9.52)

(9.53)

(9.54)

It is possible to develop the term for the partial pressure of each constituent from statement (b) of the Gibbs-Dalton laws. Then, for constituent a occupying the total volume of the mixture at the pressure and temperature of the mixture,

paV = na% T

But, for the mixture as a whole (eqn (9.2))

pV=n%T

Hence, by dividing eqn (9.50) by eqn (9.2)

(9.55)

giving

p a = xap , and in general pi = xip (9.56)

and, also n n n

C P i = X i P = P c xi = P (9.57) ; = I i = l i = 1

Often mixtures are analysed on a volumetric basis, and from the volumetric results it is possible to obtain the partial pressures and mole fractions of the components. Normally

Page 190: Advanced Thermodynamics for Engineers

174 Thermodynamic properties of ideal gases and ideal gas mixtures

volumetric analyses are performed at constant pressure and temperature. Then, considering the ith component,

ni%T v=- P

and the total volume of the mixture, V , is

n%T V=-- P

Hence, from eqns (9.58) and (9.59)

= x; V n , = - - V n

(9.58)

(9.59)

(9.60)

If the gas composition is given in volume percentage of the mixture, the mole fraction is

V(%) ' 100

x. = -

and the partial pressure is

V(%) Pi = - P

100

(9.61)

(9.62)

The terms relating to the energy of a mixture can be evaluated from statement (c) of the Gibbs-Dalton laws. If e, = molar internal energy, and h, = molar enthalpy, then for the mixture

n

E = ne, = 1 nie,,i

and n

H = nh,,, = 1 nih,,,i i = 1

If the molar internal energy and molar enthalpy of the mixture are required then

1 " n

e, = - nie,,,,; = C xiem,, n i = l i = 1

and

1 " n

= - 1 nik , i = C x i k . i n i = l i = 1

(9.63)

(9.64)

(9.65)

(9.66)

Neglecting motion, gravity, electricity, magnetism and capillary effects, then e, = u,, and hence

(9.67) urn = C xiurn, i

Page 191: Advanced Thermodynamics for Engineers

The definition of enthalpy for an ideal gas is

h,=u,+%T

Thus, for the mixture

hm=C(xi(Urn + R T ) i > = C X ~ U , , ~ + = C x ~ u , , ~ + %T

xi%T = C x;u,,; + RT C xi

Entropy of mixtures 175

(9.68)

(9.69)

9.4.4 SPECIFIC HEATS OF MIXTURES

Statement (c) of the Gibbs-Dalton law and the above expressions show that

l.4, = c xiu,,i and

hrn = C xihrn,i By definition, the specific heats are

du dh

' dT dT c = - and cP=-.

Thus

and, similarly

cp. rn = C xi (cp, rn)i

(9.70)

(9.71)

9.5 Entropy of mixtures

Consider a mixture of two ideal gases, a and b. The entropy of the mixture is equal to the sum of the entropies which each component of the mixture would have i f i r alone occupied the whole volume of the mixture at the same temperature. This concept can present some difficulty because it means that when gases are separate but all at the same temperature and pressure they have a lower entropy than when they are mixed together in a volume equal to the sums of their previous volumes; the situation can be envisaged from Fig 9.2. This can be analysed by considering the mixing process in the following way:

(i) Gases a and b, at the same pressure, are contained in a control volume but prevented from mixing by an impermeable membrane.

(ii) The membrane is then removed, and the components mix due to diffusion, i.e. due to the concentration gradient there will be a net migration of molecules from their original volumes. This means that the probability of finding a particular molecule at any particular point in the volume is decreased, and the system is in a less ordered state. This change in probability can be related to an increase in entropy by statistical thermodynamics. Hence, qualitatively, it can be expected that mixing will give rise to an increase in entropy.

Page 192: Advanced Thermodynamics for Engineers

176 Thermodynamic properties of ideal gases and ideal gas mixtures

Fig. 9.2 Two gases at the same pressure, p , contained in an insulated container and separated by a membrane; p , = p b = p

Considering the mixing process from a macroscopic viewpoint, when the membrane is broken the pressure is unaffected but the partial pressures of the individual components are decreased.

The expression for the entropy of a gas is (eqn (9.20))

P

Po s, = s,(T) - 3 In - + SO,,

where s,(T) is a function of T alone. Hence, considering the pressure term, which is actually the partial pressure of a component, a decrease in the partial pressure will cause an increase in the entropy of the gas. Equation (9.20) can be simplified by writing the pressure ratio as p , = p i p o , giving

s, = s,(T) - % In p , + so,,

Now consider the change from an analytical viewpoint. Consider the two gases a and b. The total amount of substance in the gases is

(9.72) nT = nu + nb

The entropies of the gases before mixing are

S m 0 = = n a [ s m o ( T ) - 3 P I , + s o , m 0 I (9.73)

S m , = n b S m b = n b [ s m b ( T ) - % In P r b + SO,rnbl (9.74)

and

giving the sum of the entropies before mixing (Le. separated) as

(SmT)sep = (nTS,T)Xp = nT[xasmd(T) + XbSmb(T) + XaSO,m, + XbSO,mb - m ( X a In Pr, + xb In P r b ) ]

(9.75)

Since the gases are separated and both are at the pressure, p, , then

(SmT)sep = nT[xasm0(T) + XbSmb(T) + XaSO,rn, + X b S O , m b - 3 In P I (9.76)

Hence, the molar entropy, when the gases are separate, is

(SmT)sep = x a s m , ( T ) + X b S m b ( T ) + xaSO,m, + XbSO,rnb - % In p = C xismz + C x,sO.,, - 3 In p (9.77)

Page 193: Advanced Thermodynamics for Engineers

Entropy of mixtures 177

The entropy of the gases after mixing is

(SmT)m,x = nTsmT = nsm = nosma + n b S q , (9.78)

where smo and smb are the molar entropies if each component occupied the whole volume. Thus, substituting for s, from eqn (9.20a) gives

(nTSmT)mix = n o s m a ( T ) + nbsmb(T) - 3 ( n , In p r , + nb In P r b ) + naSO,rn, -t nbSO,rnb

(SmT)mix = xasrno(T) + X b S r n b ( T ) - m ( X a In Pr, + x b In pr,) + XosO,rn, + XbSO,mb

(9'79)

(9.80)

where pr, = p I / p o , and p , = partial pressure of constituent i. Equation (9.80) may be written

(SmTIrnix = C xisrn,(T) + C XzsO,rn, - 3 C 1, In ~ r , (9.81)

Now, consider the final term in eqn (9.81), based on two constituents to simplify the mathematics

x , In P I , + xb In P r b = xa h ( x a P r > + xb ln(xbpr)

= x , In x , + x , In p r + xb In xb + x b In p r = C x , In x , + In pr C x ,

1 = C x, In x, + In pr = In pr - C xi In -

X i

Hence

1 ( s , , , ~ ) ~ ~ = C x,s,,(i") + C x,so,,, + 3

Considering the terms on the right-hand side of eqn (9.83):

(9.82)

(9.83)

1 st and 2nd terms 3rd term

summation of entropies before mixing, ( SnlT)sep change of entropy due to mixing; this is due to a change ofpartial pressures when mixed

4th term pressure term

The change of entropy due to mixing is given by

AS = (Srn,)rnix - (SrnrIFep

which is the difference between eqns (9.83) and (9.77). This gives

1 AS = 3 c x, In -

X ,

(9.84

(9.85

The numerical value of AS must be positive because x , is always less than unity. Equation (9.85) shows that there is an entropy increase due to mixing, and this is caused

by the reduction in the order of the molecules. Before mixing it is possible to go into one side of the container and guarantee taking a particular molecule, because only molecules of a are in the left-hand container, and only molecules of b are in the right-hand container. After mixing it is not possible to know whether the molecule obtained will be of a or 6 : the order of the system has been reduced.

I

Page 194: Advanced Thermodynamics for Engineers

178 Thermodynamic properties of ideal gases and ideal gas mixtures

What happens if both a and b in Fig 9.2 are the same gas?

Superficially it might be imagined that there will still be an increase of entropy on mixing. However, since a and b are now the same there is no change in partial pressure due to mixing. This means that the molar fraction is unaltered and

A S = O (9.86)

9.6 Concluding remarks

A detailed study of ideal gases and ideal gas mixtures has been undertaken in preparation for later chapters. Equations have been developed for all properties, and enthalpy coefficients have been introduced for nine commonly encountered gases. Tables of gas properties have been presented for most gases occurring in combustion calculations.

Equations for gas mixtures have been developed and the effects of mixing on entropy and Gibbs energy have been shown.

PROBLEMS

Assume that air consists of 79% N, and 21% 0, by volume.

A closed vessel of 0.1 m3 capacity contains a mixture of methane ( C h ) and air, the air being 20% in excess of that required for chemically correct combustion. The pressure and temperature in the vessel before combustion are, respectively, 3 bar and 100°C. Determine: (a) the individual partial pressures and the weights of methane, nitrogen and

oxygen present before combustion; (b) the individual partial pressures of the burnt products, on the assumption that

these are cooled to 100°C without change of volume and that all the vapour produced by combustion is condensed.

[(a) 0.2415, 2.179,0.579 bar; 0.01868,0.2951,0.08969 kg (b) 2.179, 0.2415,0.0966 bar]

1

2 An engine runs on a rich mixture of methyl and ethyl alcohol and air. At a pressure of 1 bar and 10°C the fuel is completely vaporised. Calculate the air-fuel ratio by volume under these conditions, and the percentage of ethyl alcohol in the fuel by weight. If the total pressure of the exhaust gas is 1 bar, calculate the dew point of the water vapour in the exhaust and the percentage by volume of carbon monoxide in the dry exhaust gas, assuming all the hydrogen in the fuel forms water vapour.

Vapour pressures at 10°C: methyl alcohol (CH,OH), 0.0745 bar; and ethyl alcohol (C&OH), 0.310 bar.

[63"C; 4.15%]

3 An engine working on the constant volume (Otto) cycle has a compression ratio of 6.5 to 1, and the compression follows the law P V ' . ~ = C, the initial pressure and temperature being 1 bar and 40°C. The specific heats at constant pressure and constant volume throughout compression and combustion are 0.96 + 0.00002T H/kg K and 0.67 + 0.00002 T H/kg K respectively, where T is in K.

Page 195: Advanced Thermodynamics for Engineers

Problems 179

Find (a) the change in entropy during compression; (b) the heat rejected per unit mass during compression; (c) the heat rejected per unit mass during combustion if the maximum pressure is

43 bar and the energy liberated by the combustion is 2150 kJ/kg of air. [(a) -0.1621 kJ/kg K; (b) -67.8 kJ/kg; (c) -1090.6 kJ/kg]

4 A compression-ignition engine runs on a fuel of the following analysis by weight: carbon 84%, hydrogen 16%. If the pressure at the end of combustion is 55 bar, the volume ratio of expansion is 15 : 1, and the pressure and temperature at the end of expansion are 1.75 bar and 600°C respectively, calculate (a) the variable specific heat at constant volume for the products of combustion; and (b) the change in entropy during the expansion stroke per kmol.

The expansion follows the law pv" = C and there is 60% excess air. The specific heats at constant volume in kJ/kmol K between 600°C and 2400°C are:

0 2 32 + 0.0025 T ; N, 29 + 0.0025 T H,O 34 + 0.08T; CO, 50 + 0.0067T

The water vapour (H,O) may be considered to act as a perfect gas. [(a) 22.1 + 0.0184T; (b) - 1 1.42 W/kmol K]

5 The exhaust gases of a compression-ignition engine are to be used to drive an exhaust gas turbocharger. Estimate the mean pressure ratio of expansion and the isentropic enthalpy drop per kmol of gas in the turbine if the mean exhaust temperature is 600°C and the isentropic temperature drop is 100°C. The composition of the exhaust gas by volume is CO,, 8%; H,O, 9.1%; O,, 7.5%; N,, 75.4%. The specific heats at constant volume in kJ/kmol K are:

0, 32 + 0.0025 T ; N2 29 + 0.0025T H,O 34+0.08T CO, 50 + 0.00677'

The water vapour (H,O) may be considered to act as a perfect gas. [1.946; -4547.11

6 (a) An amount of substance equal to 2kmol of an ideal gas at temperature T and pressure p is contained in a compartment. In an adjacent compartment is an amount of substance equal to 1 kmol of an ideal gas at temperature 2T and pressure p . The gases mix adiabatically but do not react chemically when the partition separating the compartments is withdrawn. Show that, as a result of the mixing process, the entropy increases by

4 K - 1 27

provided that the gases are different and that IC, the ratio of specific heats, is the same for both gases and remains a constant in the temperature range T to 2T. (b) What would be the entropy change if the gases being mixed were of the same species?

Page 196: Advanced Thermodynamics for Engineers

180 Thermodynamic properties of ideal gases and ideal gas mixtures

7 The exhaust gas from a two-stroke cycle compression-ignition engine is exhausted at an elevated pressure into a large chamber. The gas from the chamber is subsequently expanded in a turbine. If the mean temperature in the chamber is 811 K and the pressure ratio of expansion in the turbine is 4 : 1 , calculate the isentropic enthalpy drop in the turbine per unit mass of gas.

[256.68]

8 The following data refer to an analysis of a dual combustion cycle with a gas having specific heats varying linearly with temperature.

The pressure and temperature of the gas at the end of compression are 31 bar and 227°C respectively; the maximum pressure achieved during the cycle is 62 bar, while the maximum temperature achieved is 1700°C. The temperature at the end of the expansion stroke is 1240°C. The increases in entropy during constant volume and constant pressure combustion are 0.882 and 1.450 kJ/kg K respectively. Assuming that the fluid behaves as an ideal gas of molecular weight m,=30.5, calculate the equations for the specific heats, and also the expansion ratio if that process is isentropic.

[0.6747 + 0.0008297'; 0.94735 + 0.000829T; 7.8141

9 Distinguish between an ideal and a perfect gas and show that in both cases the specific entropy, s, is given by

s = s , + I;; - - .I.(:)

Two streams of perfect gases, A and B, mix adiabatically at constant pressure and without chemical change to form a third stream. The molar specific heat at constant pressure, c ~ , ~ , of the gas in stream A is equal to that in stream B. Stream A flows at M kmol/s and is at a temperature T,, while stream B flows at 1 kmol/s and is at a temperature nT,. Assuming that the gases A and B are different, show that the rate of entropy increase is

cp In [ - 1, (::;).+'I - - . h[; (A)..'] How is the above expression modified if the gases A and B are the same?

(a) (b)

For the case n = 1, evaluate the rate of entropy increase

when different gases mix, and when the gas in each stream is the same.

[c,, In[ -!- (E)."]; (a) 3 h[ (")""I: (b) O] n M + l M M + l

10 A jet engine bums a weak mixture of octane (C,H,,) and air, with an equivalence ratio, @ = 2. The products of combustion, in which dissociation may be neglected, enter the nozzle with negligible velocity at a temperature of 1000 K. The gases, which may be considered to be ideal, leave the nozzle at the atmospheric pressure of

Page 197: Advanced Thermodynamics for Engineers

Problems 181

1.013 bar with an exit velocity of 500 m/s. The nozzle may be considered to be adiabatic and frictionless.

Determine:

(a)

(b) (c) (d)

Specific heat at constant pressure, cp,,,, in J/kmol K, with Tin K:

the specific heat at constant pressure, cp, of the products as a function of temperature; the molecular weight, m,, of the products; the temperature of the products at the nozzle exit; the pressure of the products at the nozzle inlet.

COZ HZO 0 2

N2

cp,,, = 21 x 10’ + 34.OT cp,,, = 33 x lo3 + 8.3T cp,,, = 28 x lo3 + 6.4T cp,,, = 29 x io3 + 3.4T

[0.9986 x lo3 + 0.2105T; 28.71; 895.7; 1.591 bar]

11 The products of combustion of a jet engine have a molecular weight, m,, of 30 and a molar specific heat at constant pressure given by cp,,, = 3.3 x lo4 + 15T J/kmol K where T is the gas temperature in Kelvin. When the jet pipe stagnation temperature is 1200 K the gases leave the nozzle at a relative speed of 600 m/s. Evaluate the static temperature of the gas at the nozzle exit and estimate the total to static pressure ratio across the nozzle. Assume that the products of combustion behave as an ideal gas and that the flow is isentropic.

In a frictional nozzle producing the mean outlet speed from the same inlet gas temperature, what would be the effect on (a) (b)

the mean outlet static temperature, and the total to static pressure ratio?

[1092; 1.732; (a) unaffected; (b) increased]

Page 198: Advanced Thermodynamics for Engineers

10 Thermodynamics of Combustion

Combustion is an oxidation process and is usually exothermic (i.e. releases the chemical (or bond) energy contained in a fuel as thermal energy). The most common combustion processes encountered in engineering are those which convert a hydrocarbon fuel (which might range from pure hydrogen to almost pure carbon, e.g. coal) into carbon dioxide and water. This combustion is usually performed using air because it is freely available, although other oxidants can be used in special circumstances, e.g. rocket motors. The theory that will be developed here will be applicable to any mixture of fuel and oxidant and any ratio of components in the products; however, it will be described in terms of commonly available hydrocarbon fuels of the type used in combustion engines or boilers.

The simplest description of combustion is of a process that converts the reactants available at the beginning of combustion into producrs at the end of the process. This model presupposes that combustion is a process that can take place in only one direction and it ignores the true statistical nature of chemical change. Combustion is the combination of various atoms and molecules, and takes place when they are close enough to interact, but there is also the possibility of atoms which have previously joined together to make a product molecule separating to form reactants again. The whole mixture is really taking part in a molecular ‘barn dance’ and the tempo of the dance is controlled by the temperature of the mixture. The process of molecular breakdown is referred to as dissociation; this will be introduced in Chapter 12. In reality a true combustion process is even more complex than this because the actual rate at which the reactions can occur is finite (even if extremely fast). This rate is the basic cause of some of the pollutants produced by engines, particularly NO,. In fact, in most combustion processes the situation is even more complex because there is an additional factor affecting combustion, which is related to the rate at which the fuel and air can mix. These ideas will be introduced in Chapter 15. Hence, the approach to combustion in this chapter is a simplified one but, in reality, it gives a reasonable assessment of what would be expected under good combustion conditions. It cannot really be used to assess emissions levels but it can be extended to this simply by the introduction of additional equations: the basic approach is still valid.

The manner in which combustion takes place is governed by the detailed design of the combustion system. The various different types of combustion process are listed in Table 10.1, and some examples are given of where the processes might be found. There is an interdependence between thermodynamics and fluid mechanics in combustion, and this interaction is the subject of current research. This book will concentrate on the thermo-

Page 199: Advanced Thermodynamics for Engineers

Thermodynamics of combustion 183

dynamics of combustion, both in equilibrium and non-equilibrium states. The first part of the treatment of combustion will be based on equilibrium thermodynamics, and will cover combustion processes both with and without dissociation. It will be found that equilibrium thermodynamics enables a large number of calculations to be performed but, even with dissociation included, it does not allow the calculation of pollutants, the production of which are controlled both by mixing rates (fluid mechanics) and reaction rates (thermodynamics).

Table 10.1 Factors affecting combustion processes

Conditions of combustion Classification Examples

Time dependence

Spatial dependence

Mixing of initial reactants

Flow

Phase of reactants

Reaction sites

Reaction rate

Convection conditions

Compressibility

Speed of combustion

steady unsteady

zero-dimensional

one-dimensional

two-dimensional three-dimensional

premixed non-premixed

laminar turbulent

single

multiphase

homogeneous heterogeneous

equilibrium chemistry (infinite rate)

finite rate

natural

forced

incompressible compressible

deflagration detonation

gas turbine combustion chamber, boilers petrol engine, diesel engine

only used for modelling purposes, well- stirred reactors approximated in pipe flows, flat flame burners axisymmetric flames, e.g. Bunsen burner general combustion

petrol, or spark ignition, engine diesel engine, gas turbine combustion chamber

special cases for measuring flame speed most real engine cases, boilers

spark-ignited gas engines, petrol engines with fuel completely evaporated; gas-fired boilers diesel engines, gas turbines, coal- and oil- fired boilers

spark-ignition engines diesel engines, gas turbines, coal fired boilers

approached by some processes in which the combustion period is long compared with the reaction rate all real processes: causes many pollutant emissions

Bunsen flame, gas cooker, central heating boiler gas turbine combustion chamber, large boilers

free flames engine flames

most normal combustion processes ‘knock’ in spark ignition engines, explosions

Page 200: Advanced Thermodynamics for Engineers

I84 Thermodynamics of combustion

10.1 Simple chemistry

Combustion is a chemical reaction and hence a knowledge of basic chemistry is required before it can be analysed. An extremely simple reaction can be written as

2 c o + 0, (j 2c0 , (10.1)

This basically means that two molecules of carbon monoxide (CO) will combine with one molecule of oxygen (0,) to create two molecules of carbon dioxide (CO,). Both CO and 0, are diatomic gases, whereas CO, is a triatomic gas. Equation (10.1) also indicates that two molecules of CO, will always break down into two molecules of CO and one molecule of 0,; this is signified by the symbol e which indicates that the processes can go in both directions. It is conventional to refer to the mixture to the left of the arrows as the reactants and that to the right as the products; this is because exothermic combustion (i.e. in which energy is released by the process) would require CO and 0, to combine to give CO,. Not all reactions are exothermic and the formation of NO during dissociation occurring in an internal combustion (i.c.) engine is actually endothermic.

It should be noted from the combustion eqn (10.1) that three molecules of reactants combine to produce two molecules of products, hence there is not necessarily a balance in the number of molecules on either side of a chemical reaction. However, there is a balance in the number of atoms of each constituent in the equation and so mass is conserved.

10.1.1 FUELS

Hydrocarbon fuels are rarely single-component in nature due to the methods of formation of the raw material and its extraction from the ground. A typical barrel of crude oil contains a range of hydrocarbons, and these are separated at a refinery; the oil might produce the constituents defined in Fig 10.1. None of the products of the refinery is a single chemical compound, but each is a mixture of compounds, the constituents of which depend on the source of the fuel.

Light distillates (chemical feedstock) / 1

Fuel to run refinery

Heavy fuel, or

A

Heavy fuel, or

A

Middle distillates (gas oil. Kerosene (paraffin, diesel, heating oil)

Gases (butane, propane)

aviation fuel)

Fig. 10.1 Typical constituents of a barrel of crude oil

Page 201: Advanced Thermodynamics for Engineers

Combustion of simple hydrocarbons fuels 185

One fuel which approaches single-component composition is ‘natural gas’, which consists largely of methane (CH,). Methane is the simplest member of a family of hydrocarbons referred to as paraffins or, more recently, alkanes which have a general formula C,,HZn+,. The lower alkanes are methane (CH,), ethane (C,H,), propane (C,H,) and butane (C,H,,) etc. Two other alkanes that occur in discussion of liquid fuels are heptane (C,H,,) and octane (CEHIE). The alkanes are referred to as saturated hydrocarbons because it is not physically possible to add more hydrogen atoms to them. However, it is possible to find hydrocarbons with less than 2n + 2 hydrogen atoms and these are referred to as unsaturated hydrocarbons. A simple unsaturated hydrocarbon is acetylene (C,H,), which belongs to a chemical family called alkenes. Some fuels contain other constituents in addition to carbon and hydrogen. For example, the alcohols contain oxygen in the form of an OH radical. The chemical symbol for methanol is CH,OH, and that for ethanol is C,H,OH; these are the alcohol equivalents of methane and ethane.

Often fuels are described by a mass analysis which defines the proportion by mass of the carbon and hydrogen, e.g. a typical hydrocarbon fuel might be defined as 87% C and 13% H without specifying the actual components of the liquid. Solid fuels, such as various coals, have a much higher carbon/hydrogen ratio but contain other constituents including oxygen and ash.

The molecular weights (or relative molecular masses) of fuels can be evaluated by adding together the molecular (or atomic) weights of their constituents. Three examples are given below:

Methane (CH,)

Octane (CEH18)

Methanol (CH,OH)

(m,),, = 12 + 4 x 1 = 16

(mw)CsH,s = 8 x 12 + 18 x 1 = 114

(mw)CH30H = 12 + 3 x 1 + 16 + 1 = 32

10.2 Combustion of simple hydrocarbon fuels

The combustion of a hydrocarbon fuel takes place according to the constraints of chemistry. The combustion of methane with oxygen is defined by

CH, + 2 Q 3 C Q + 2H,O 1 kmol 2 kmol 1 kmol 2 kmol 1 2 + 4 2 x 3 2 12x32 2 x ( 2 + 1 6 ) 16kg 64kg 44 kg 36 kg

(10.2)

In this particular case there is both a molar balance and a mass balance: the latter is essential but the former is not. Usually combustion takes place between a fuel and air (a mixture of oxygen and nitrogen). It is normal to assume, at this level, that the nitrogen is an inert gas and takes no part in the process. Combustion of methane with air is given by

79

21 a CO, + 2H2O + 2 ~ - ” , (10.3)

1 kmol 9.52 kmol 1 kmol 2 kmol 7.52 kmol 1 2 + 4 2x(32+105.2) 12+32 2x(2+16) 7 . 5 2 ~ 2 8 16 kg 274.4 kg 44 kg 36 kg 210.67 kg

Page 202: Advanced Thermodynamics for Engineers

186 Thermodynamics of combustion

10.2.1 STOICHIOMETRY

There is a clearly defined, and fixed, ratio of the masses of air and fuel that will result in complete combustion of the fuel. This mixture is known as a stoichiometric one and the ratio is referred to as the stoichiometric air- fuel ratio. The stoichiometric air-fuel ratio, for methane can be evaluated from the chemical equation (eqn 10.3). This gives

mass of air

mass of fuel 16

2 x (32 + 105.33) = 17.17 - & . = -

StOlC

This means that to obtain complete combustion of 1 kg CH, it is necessary to provide 17.17 kg of air. If the quantity of air is less than 17.17 kg then complete combustion will not occur and the mixture is known as rich. If the quantity of air is greater than that required by the stoichiometric ratio then the mixture is weak.

10.2.2 COMBUSTION WITH WEAK MIXTURES

A weak mixture occurs when the quantity of air available for combustion is greater than the chemically correct quantity for complete oxidation of the fuel; this means that there is excess air available. In this simple analysis, neglecting reaction rates and dissociation etc, this excess air passes through the process without taking part in it. However, even though it does not react chemically, it has an effect on the combustion process simply because it lowers the temperatures achieved due to its capacity to absorb energy. The equation for combustion of a weak mixture is

2 7.52

$J CH, + - (0, + 3.76Nz)

where @ is called the equivalence ratio, and

actual fuel-air ratio

stoichiometric fuel-air ratio

stoichiometric air-fuel ratio

actual air-fuel ratio @ =

For a weak mixture q5 is less than unity. Consider a weak mixture with

CH, + 2.5(0, + 3.76N2) 3 CO, + 2H,O + 0.50, + 9.4Nz

NZ ( 10.4)

(10.5)

c$ = 0.8; then

(10.6)

10.2.3 COMBUSTION WITH RICH MIXTURES

A rich mixture occurs when the quantity of air available is less than the stoichiometric quantity; this means that there is not sufficient air to bum the fuel. In this simplified approach it is assumed that the hydrogen combines preferentially with the oxygen and the carbon does not have sufficient oxygen to be completely burned to carbon dioxide; this results in partial oxidation of part of the carbon to carbon monoxide. It will be shown in Chapter 12 that the equilibrium equations, which control the way in which the hydro- carbon fuel oxidizes, govern the proportions of oxygen taken by the carbon and hydrogen of the fuel and that the approximation of preferential combination of oxygen and

Page 203: Advanced Thermodynamics for Engineers

Heats of formation and heats of reaction 187

hydrogen is a reasonable one. In this case, to define a rich mixture, @ is greater than unity. Then

7.52 co+- N2 2 ( :@)co2 + 2H20 + 4(@ - 1) CH4 + - ( 0 2 + 3.76N2) * ___ @ @ @

(10.7)

If the equivalence ratio is 1.2, then eqn (10.7) is

CH, + 1.667(0, + 3.76N2) 3 0.333C0, + 2H,O + 0.667CO + 6.267N2 (10.8)

It is quite obvious that operating the combustion on rich mixtures results in the production of carbon monoxide (CO), an extremely toxic gas. For this reason it is now not acceptable to operate combustion systems with rich mixtures. Note that eqn (10.7) cannot be used with values of @ >4/3, otherwise the amount of CO, becomes negative. At this stage it must be assumed that the carbon is converted to carbon monoxide and carbon. The resulting equation is

2 CH4+-(02+3.76N2)*2H20+

@ Equation (10.9) is a very hypothetical one because during combustion extensive dissociation occurs and h s liberates oxygen by breaking down the water molecules; this oxygen is then available to create carbon monoxide and carbon dioxide rather than carbon molecules.

In reality it is also possible to produce pollutants even when the mixture is weaker than stoichiometric, simply due to poor mixing of fuel and air, quenching of flames on cold cylinder or boiler walls, trapping of the mixture in crevices (fluid mechanics effects) and also due to thermodynamic limitations in the process.

10.3 Heats of formation and heats of reaction

Combustion of fuels takes place in either a closed system or an open system. The relevant property of the fuel to be considered is the internal energy or enthalpy, respectively, of formation or reaction. In a naive manner it is often considered that combustion is a process of energy addition to the system. This is not true because the energy released during a combustion process is already contained in the reactants, in the form of the chemical energy of the fuel (see Chapter 11). Hence it is possible to talk of adiabatic combustion as a process in which no energy (heat) is transferred to, or from, the system - the temperature of the system increases because of a rearrangement of the chemical bonds in the fuel and oxidant.

Mechanical engineers are usually concerned with the combustion of hydrocarbon fuels, such as petrol, diesel oil or methane. These fuels are commonly used because of their ready availability (at present) and high energy density in terms of both mass and volume. The combustion normally takes place in the presence of air. In some other applications, e.g. space craft, rockets, etc, fuels which are not hydrocarbons are burned in the presence of other oxidants; these will not be considered here.

Hydrocarbon fuels are stable compounds of carbon and hydrogen which have been formed through the decomposition of animal and vegetable matter over many millennia. It is also possible to synthesise hydrocarbons by a number of processes in which hydrogen is

Page 204: Advanced Thermodynamics for Engineers

188 Thermodynamics of combustion

added to a carbon-rich fuel. The South African Sasol plant uses the Lurgi and Fischer-Tropsch processes to convert coal from a solid fuel to a liquid one. The chemistry of fuels is considered in Chapter 11.

10.4 Application of the energy equation to the combustion process - a macroscopic approach

Equations (10.3) to (10.6) show that combustion can take place at various air-fuel ratios, and it is necessary to be able to account for the effect of mixture strength on the combustior. process, especially the temperature rise that will be achieved. It is also necessary to be able to account for the different fuel composition: not all fuels will release the same quantity of energy per unit mass and hence it is required to characterise fuels by some capacity to release chemical energy in a thermal form. Both of these effects obey the First Law of Thermodynamics, i.e. the energy equation.

10.4.1

It was shown previously that the internal energies and enthalpies of ideal gases are functions of temperature alone ( cp and c, might still be functions of temperature). This means that the internal energy and enthalpy can be represented on U-T and H-T diagrams. It is then possible to draw a U-T or H-T line for both reactants and products (Fig 10.2). The reactants will be basically diatomic gases (neglecting the effect of the fuel) whereas the products will be a mixture of diatomic and triatomic gases - see eqn (10.3).

INTERNAL ENERGIES AND ENTHALPIES OF IDEAL GASES

Temperature, T

Fig. 10.2 Enthalpy (or internal energy) of reactants and products

The next question which arises is: what is the spacing between the reactants and products lines? This spacing represents the energy that can be released by the fuel.

10.4.2

The energy contained in the fuel can also be assessed by burning it under a specified condition; this energy is referred to as the heat of reaction of the fuel. The heat of reaction

HE.4TS OF REACTION AND FORMATION

Page 205: Advanced Thermodynamics for Engineers

Application of the energy equation to the combustion process 189

for a fuel is dependent on the process by which it is measured. If it is measured by a constant volume process in a combustion bomb then the internal energy of reaction is obtained. If it is measured in a constant pressure device then the enthalpy of reaction is obtained. It is more normal to measure the enthalpy of reaction because it is much easier to achieve a constant pressure process. The enthalpy of reaction of a fuel can be evaluated by burning the fuel in a stream of air, and measuring the quantity of energy that must be removed to achieve equal reactant and product temperatures (see Fig 10.3).

Control surface

I T,= T,

e- QP

Fig. 10.3 Constant pressure measurement of enthalpy reaction

Applying the steady flow energy equation

( : ) ( : ) Q - W s = liz, he + - + gz, - ri2; hi + - + gz, (10.10)

and neglecting the kinetic and potential energy terms, then

( Q P h = - ( H R ) T = nP(hP)T - n R ( h d ~ (1G.11)

where n denotes the amount of substance in either the products or reactants; this is identical to the term n which was used for the amount of substance in Chapter 9. The suffix T defines the temperature at which the enthalpy of reaction was measured. is a function of this temperature and normally it is evaluated at a standard temperature of 25°C (298 K). When is evaluated at a standardised temperature it will be denoted by the symbol (ep),. Most values of Qp that are used in combustion calculations are the (ep), ones. (In a similar way, (Q,), will be used for internal energy of reaction at the standard temperature.) The sign of Qp is negative for fuels because heat must be transferred from the ‘calorimeter’ to achieve equal temperatures for the reactants and products (it is positive for some reactions, meaning that heat has to be transferred to the calorimeter to maintain constant temperatures). The value of the constant volume heat of reaction, the internal energy of reaction, (e,),, can be calculated from (ep), as shown below, or measured using a constant volume combustion ‘bomb’; again (Q,), has a negative value. (ep), and (Q,), are shown in Figs 10.4(a) and (b) respectively. The term calorific value of the fuel was used in the past to define the ‘heating’ value of the fuel: this is actually the negative value of the heat of reaction, and is usually a positive number. It is usually associated with analyses in which ‘heat’ is added to a system during the combustion process, e.g. the air standard cycles.

Applying the first law for a closed system to constant volume combustion gives

(10.12)

Page 206: Advanced Thermodynamics for Engineers

190 Thermodynamics of combustion

If both the products and reactants are ideal gases then h = J c ~ , ~ dT, and u = J cy,,, dT, which can be evaluated from the polynomial expressions derived in Chapter 9. Thus

( Q p I s - ( Q v I s = ~ P ( ~ P ) T - ~ R ( ~ R ) T - ( ~ P ( ~ P ) T - n R ( u R ) T l

= np( (hP)T- (up)TJ - nRI (hR)T- (uR)T} = %T(np - nR) (10.13)

This result is quite logical because the definitions of (Qp)s and (e,), require that Tp and TR are equal. Hence the constant pressure and constant volume processes are identical if the amounts of substance in the products and the reactants are equal. If the amounts of substance change during the reaction then the processes cease to be identical and, in the case of a combustion bomb, a piston would have to move to maintain the conditions. The movement of the piston is work equal to %T(np - nR).

It is also possible to relate the quantity of energy that is chemically bound up in the fuel to a value at absolute zero of temperature. These values are denoted as -AHo and -AUo and will be returned to later.

10.4.3 HEAT OF FORMATION - HESS' LAW

The heat of formation of a compound is the quantity of energy absorbed (or released) during its formation from its elements (the end pressures and temperatures being maintained equal).

For example, if CO, is formed from carbon and oxygen by the reaction

c + 0, -+ co, (10.14)

then in a constant pressure steady flow process with equal temperature end states the reaction results in heat transfer of ( given by

<ep) T = HP - H R

= -394 MJ/kmol

Reactants / x e

T, Temperature, T

(10.15)

(10.16)

Reactants

/*

, 'Products 0 ,

0 , /

Temperature, T

(b)

Fig. 10.4 Internal energy and enthalpy of reaction depicted on (a) internal energy-temperature and (b) enthalpy -temperature diagrams

Page 207: Advanced Thermodynamics for Engineers

Application of the energy equation to the combustion process 191

If a slightly different reaction is performed giving the same end product, e.g.

1

2 co + - 0 2 - co, (10.17)

then it is not possible to use the same simple approach because the reactants are a mixture of elements and compounds. However, Hess’ law can be used to resolve this problem. This states that:

(a) if a reaction at constant pressure or constant volume is carried out in stages, the algebraic sum of the amounts of heat evolved in the separate stages is equal to the total evolution of heat when the reaction occurs directly;

or

(b) the heat liberated by a reaction is independent of the path of the reaction between the initial and final states.

Both of these are simply statements of the law of energy conservation and the definition of properties. However, this allows complex reactions to be built up from elemental ones. For example, the reaction

1

2 co + - 0, - CO,

can be sub-divided into two different reactions:

1

2 c + - 0, + co

1

2 co + - 0, - CO,

(10.18)

(1 0.19a)

(10.19b)

The heat of formation of CO may be evaluated by reaction (10.19a) and then used in reaction (10.19b) to give the heat of reaction of that process. From experiment the heat of formation of carbon monoxide (CO) is - 112 MJ/kmol and hence, for reaction (10.19b), the energy released is

(Q,), = Hp- HR= -394- (-110.5)= -283 MJ/kmol (10.20)

Hence the heat of reaction for CO + io2 -.+ CO, is - 283 MJ/kmol. The heats of formation of any compounds can be evaluated by building up simple

reactions, having first designated the heats of formation of elements as zero. This enables the enthalpy of formation of various compounds to be built up from component reactions. Enthalpies of formation are shown in Table 10.2.

These enthalpies of formation can be used to evaluate heats of reaction of more complex molecules, e.g. for methane (CH,) the equation is

CH, + 20, = CO, + 2H,O (g)

giving

(12,125 = (AH,)co,,g) + 2(AHfh2o(g) - (AHf)c,,g, = -802 279 k.J/kmol

This gives an enthalpy of reaction per kg of CH, of -50 142 kJ/kg.

Page 208: Advanced Thermodynamics for Engineers

192 Thermodynamics of combustion

Table 10.2 Enthalpies of formation of some common elements and compounds

Species Reaction

Oxygen, 0, Hydrogen, H, Carbon, C Carbon dioxide, CO, Carbon monoxide, CO Water vapour, H,O Water (liquid), H,O Nitric oxide (NO) Methane, CH, Ethane, C2H, Propane, C,H, Butane, C,H,, Iso-octane, C,H,, Iso-octane, C,H,, Methyl alcohol, CH,OH Methyl alcohol, CH,OH Ethyl alcohol, C,&OH Ethyl alcohol, C,H50H

c + 0,-co, c + o - c o H,+ 1/20,--2O(g) H, + 1/20,-H20(1) 1/2N, + 1/20,-NO c + 2H,-CCH,(g) 2C + 3H2--C,&(g) 3C + 4H,-Cc,H8(g) 4 c + 5H*-C4HlO(g) 8C + 9H,--+C&Mg) 8 c + 9H,-C8Hls(l) C + 2H2 + 1/2O2-CH3OH(g) C + 2H, + 1/20,-CH30H(1) 2C + 3H2 + 1/20,-+C,&OH(g) 2C + 3H, + 1/202-C2&0H(1)

gas 25"C, 1 atm gas25"C, 1 atm gas 25"C, 1 atm gas 25"C, I amm* gas 25"C, I atm* gas 25"C, I atm* liquid 25"C, 1 atm gas 25"C, 1 atm* gas 25"C, I atm* gas 25°C. 1 atm gas 25"C, 1 atm gas25"C, 1 atm gas 25"C, 1 atm liquid 25"C, 1 atm gas 25"C, 1 atm liquid 25"C, 1 atm gas 25"C, 1 atm liquid 25"C, 1 atm

0, element 0, element 0, element -393 522 - 110 529 -241 827 -285 800 +89 915 -74 897 -84 725 - 103 916 -124 817 -224 100 -259 280 -201 200 -238 600 -234 600 -277 000

* Note: the values of AHf given m the tables of gas properties are based on 0 K.

Higher and lower calorific values

The enthalpy of reaction of CH, derived above (-50 142 kJ/kg) is the negative of the lower calorific value (LCV) because it is based on gaseous water (H,O(g)) in the products. If the water in the products (exhaust) was condensed to a liquid then extra energy could be released by the process. In this case the enthalpy of reaction would be

( Q p h = (Afff)co2 + ~ ( A H ~ ) H ~ O ( , ) - (Afff)c& = -890 225 kJ/kmol

This gives an enthalpy of reaction per kg of CH, of -55 639 kJ/kg. This is the negative of the higher calorific value (HCV).

Normally the lower calorific value, or lower internal energy or enthalpy of reaction, is used in engine calculations because the water in the exhaust system is usually in the vapour phase.

10.5 Combustion processes

10.51 ADIABATIC COMBUSTION

The heats of reaction of fuels have been described in terms of isothermal processes, i.e. the temperature of the products is made equal to the temperature of the reactants. However, it is common experience that combustion is definitely not isothermal; in fact, its major characteristic is to raise the temperature of a system. How can this be depicted on the enthalpy- temperature diagram?

Page 209: Advanced Thermodynamics for Engineers

Combustion processes 193

Consider a constant pressure combustion process, as might occur in a gas turbine, in which there is no heat or work transfer (Fig 10.5(a)).

Insulation

Rssss r - - - - - - -

T R I Gasturbine: T d combustion I 2

T, Temperature, T

Fig. Adiabatic corr.-ustion L-picted on enthalpy -temperature diagram: (a) schematic of gas turbine combustion chamber; (b) enthalpy -temperature diagram of adiabatic combustion process

Applying the steady flow energy equation, eqn (lO.ll) , to the combustion chamber shown in Fig 10.5(a) gives

Hp = HR (10.21)

Hence adiabatic combustion is a process which occurs at constant enthalpy (or internal energy, in the case of combustion at constant volume), and the criterion for equilibrium is that the enthalpies at the beginning and end of the process are equal. This process can be shown on an H-T diagram (Fig 10.5(b)) and it is possible to develop from this an equation suitable for evaluating the product temperature. The sequence of reactions defining combustion are denoted by the 'cycle ABCDA'. Going clockwise around the cycle, from A, gives

- ( Q p ) s + [HR(TR)-HR(T,)I - [Hp(Tp)-Hp(Ts)I = O (10.22)

The term [HR(TR) - HR(Ts ) ] is the difference between the enthalpy of the reactants at the temperature at the start of combustion (TR) and the standardised temperature (Ts ) . The term [Hp(Tp) - H p ( T s ) ] is a similar one for the products. These terms can be written as

i = 1 i = 1

(10.23)

where i is the particular component in the reactants and q is the number of components over which the summation is made. From eqn (10.22)

HP('P) = - ( Q P ) ~ + [HR(TR) - f f R ( T s ) I + HAT, ) (1 0.24)

If the process had been a constant volume one, as in an idealised (i.e. adiabatic) internal combustion engine, then

(10.25) UP('P) = - ( Q v > s + [UR(TR) - uR(Ts) l + Up(Ts)

Page 210: Advanced Thermodynamics for Engineers

194 Thermodynamics of combustion

This possibly seems to be a complex method for evaluating a combustion process compared with the simpler heat release approach. However the advantage of this method is that it results in a true energy balance: the enthalpy of the products is always equal to the enthalpy of the reactants. Also, because it is written in terms of enthalpy, the variation of gas properties due to temperature and composition changes can be taken into account correctly. An approach such as this can be applied to more complex reactions which involve dissociation and rate kinetics: other simpler methods cannot give such accurate results.

10.5.2

The combustion process shown in Fig 10.5(b) is an adiabatic one, and the enthalpy of the products is equal before and after combustion. However, if there is heat transfer or work transfer taking place in the combustion process, as might occur during the combustion process in an internal combustion engine, then Fig 10.5(b) is modified to that shown in Fig 10.6.

COMBUSTION WITH HEAT AND WORK TRANSFER

f - AUo

L TS

Reactants / Products

Temperature afier work and heat

Temperature after work withdrawn

Temperature, T

Fig. 10.6 Combustion with heat transfer and work output

If the engine provides a work output of pdV, the temperature of the products is reduced as shown in Fig 10.6, and if there is heat transfer from the cylinder, -AQ, this reduces the temperature still further. These effects can be incorporated into the energy equation in the following way. The First Law for the process is

AQ-AW=dU= Up(Tp)- UR(TR) (10.26)

Hence

Up(Tp) = UR(TR) + AQ - AW (10.27)

This means that heat transfer away from the cylinder, as will normally happen in the case of an engine during the combustion phase, tends to reduce the internal energy of the products compared with the reactants. Also, if work is taken out of the cylinder due to the piston moving away from top dead centre, the internal energy is further reduced.

Page 211: Advanced Thermodynamics for Engineers

Examples 195

10.5.3 INCOMPLETE COMBUSTION

The value of enthalpy (or internal energy) of reaction for a fuel applies to complete combustion of that fuel to carbon dioxide and water. However, if the mixture is a rich one then there will be insufficient oxidant to convert all the fuel to CO, and H,O, and it will be assumed here that some of the carbon is converted only to CO. The chemical equations of this case are shown in Section 10.2.3. The effect of incomplete combustion on the energy released will be considered here. Incomplete combustion can be depicted as an additional line on the enthalpy-temperature diagram - the position of this line indicates how far the combustion process has progressed (see Fig 10.7). In this case the amount of energy released can be evaluated from the energies of formation, in a similar manner to that in eqn (10.8).

Products

‘ ’ Products (incomplete combustion)

(complete combustion)

T, for incomplete combustion T, for complete combustion

r

7

Temperature, T

Fig. 10.7 Enthalpy-temperature diagram for incomplete combustion including rich mixtures and dissociation

10.6 Examples

These examples quote solutions to the resolution of a calculator. This is to aid readers in identifying errors in their calculations, and does not mean the data are accurate to this level.

Example I : incomplete combustion Evaluate the energy released from constant pressure combustion of a rich mixture of methane and air: use @ = 1.2.

Solution The chemical equation is

CH, + 1.667 (0, + 3.76N2) 40 .333C02 + 2H,O + 0.667CO + 6.267N2 (10.28) The nitrogen (N,) takes no part in this reaction and does not need to be considered. The energy liberated in the reaction (ep) is the difference between the enthalpies of formation of the products and reactants.

Page 212: Advanced Thermodynamics for Engineers

196 Thermodynamics of combustion

Enthalpy of formation of reactants,

(AHf)R = (AHf),-b = -74 897 kJ/kmol

Enthalpy of formation of products,

(AHf)p = O.333(AHf),,, + 2(AHf)"*, + O.667(AHf),, = 0.333 ( -393 522) + 2( -241 827) + 0.667( - 1 10 529) = -688 420.0 kJ/km01.

Hence the energy released by this combustion process is

Qp = (AHf)p - (AHfh = -688 420.0 - (-74 897) = -613 522.0 kJ/kmol

It can be seen that this process has released only 76.5% of the energy that was available from complete combustion (-802 279 kJ/kmol). This is because the carbon has not been fully oxidised to carbon dioxide, and energy equal to 0.667 x -283 MJ/kmol CH, (equivalent to the product of the quantity of carbon monoxide and the heat of reaction for eqn (10.18)) is unavailable in the form of thermal energy. This is equivalent to 188 761 kJ/kmol C h , and is the remaining 23.5% of the energy which would be available from complete combustion. While the incomplete combustion in eqn (10.34) has occurred because of the lack of oxygen in the rich mixture, a similar effect occurs with dissociation, when the products are partially broken down into reactants.

Example 2: adiabatic combustion in an engine A stoichiometric mixture of methane (CK) and air is burned in an engine which has an effective compression ratio of 8 : 1. Calculate the conditions at the end of combustion at constant volume if the initial temperature and pressure are 27°C and 1 bar respectively. The lower calorific value of methane is 50 144 kJ/kg at 25°C. Assume the ratio of specific heats ( K ) for the compression stroke is 1.4. See Fig 10.8.

Volume

Fig. 10.8 p - V diagram for constant volume combustion

Page 213: Advanced Thermodynamics for Engineers

Examples 197

Solution

p 1 = 1 bar, T, = 300 K 1.4

p 2 = p l ( 2 ) = 1 x 81.4 = 18.379 bar

Y- 1

T2 = TI( 2) = 300 x = 689.2 K

The chemical equation is

CH, + 2(0 , + 3.76N2) - CO, + 2 H 2 0 + 7.S2N2 nR = i0.52 np= 10.52

The combustion equation is

UP(',) = - <Qv>s [UR(TR) - UR(Ts)] + Up(Ts)

The reactants' energy [ UR(TR) - UR(T,)] is

Constituent CH'l 0 2 N2

'689.2 21 883.9 15 105.9 14 587.7 '298 57 19.6 6032.0 6025.8 Difference 16 164.3 9073.9 8561.9 n 1 2 7.52 nu 16 164.3 18 146.4 64 385.5

Thus U,(TR) - UR(T,) = 9.8696 x lo4 kJ The products energy at standard temperature, Up(Ts ) , is

Constituent CO2 H2O N2

'298 7041.6 7223.3 6025.8 n 1 2 7.52 nU298 7041.6 14 446.6 45 314.0

(10.29)

(10.30)

Hence U,(Ts) = 66 802.2 kJ

terms gives Substituting these values into eqn (10.30), and converting the calorific value to molar

U,(T,) = SO 144 x 16 + 9.8696 x lo4 + 66 802.2 =9.678 x lo5 kJ

The value of specific internal energy for the mixture of products is

U,(T,) 9.678 x lo5

np 10.52 u,(T,) = ~ - - = 92000 k J / k m ~ l

Page 214: Advanced Thermodynamics for Engineers

198 Thermodynamics of combustion

An approximate value of the temperature can be obtained by assuming the exhaust products to be 70% N, and 30% CO,. Using this approximation gives

2500 76 848.7 3000 94623.3

Hence Tp is around 3000 K. Assume Tp = 3000 K and evaluate Up(Tp):

Constituent COZ HZO NZ

~ 3 o o o 137 950.4 109 612.6 76 504.6 n 1 2 7.52 nu 137 950.4 219 225.2 575 314.6

1 nu = Up(Tp) = 932 490.2 kJ

This value is less than that obtained from eqn (10.30), and hence the temperature was underestimated.

Try Tp = 3100 K.

U3lDO 143 420.0 114 194.9 79 370.9 n 1 2 7.52 nu 143 420.0 228 389.8 596 869.2

1 nu = Up(Tp) = 968 678.9 kT

Hence the value Tp is between 3000 K and 3100 K. Linear interpolation gives

967 800 - 932 490.2

968 678.9 - 932 490.2 TP = 3000 + x 1000 = 3097 K

Evaluating Up(Tp) at Tp = 3097 K gives

Constituent COZ H20 NZ

1(3097 143 255.1 114 057.5 79 285.1 n 1 2 7.52 nu 143 255.1 228 114.9 596 224

1 nu = Up(Tp) = 967 594 kJ

This is within 0.0212% of the value of Up(Tp) evaluated from eqn (10.30). The final pressure, p 3 , can be evaluated using the perfect gas law because

p3V3 = n,%T3 = np%T3

Page 215: Advanced Thermodynamics for Engineers

Examples 199

x 8 x 1 = 82.59 bar 3097 300

=-

Example 3: non-adiabatic combustion in an engine

In the previous calculation, Example 2, the combustion was assum% to be adiabatic. If 10% of the energy liberated by the fuel was lost by heat transfer, calculate the final temperature and pressure after combustion.

Solution

The effect of heat transfer is to reduce the amount of energy available to raise the temperature of the products. This can be introduced in eqn (10.30) to give

(10.31) U P V P ) = - ( Q v I s - QHT

In this case Qm= -0.l(Qv), and hence

[UR(TR) - UR(Ts)l + UdTs)

UP(TP) = -0-9(Qv>s [UR(TR) - UdTs)1 + Up(TS) (10.32)

Substituting values gives

Up(Tp) = 0.9 x 50 144 x 16 + 9.8696 x lo4 + 66 802.2 = 887 571.8 kJ

Using a similar technique to before, the value of up(Tp) is

887 571.8 10.52

UP(TP) = = 84369.9 kJ/kmol

From the previous calculation the value of Tp lies between 2500 K and 3000 K, and an estimate is

84 369.9 - 76 848.7

94 623.3 - 76 848.7 5oo = 2717 T p = 2500 +

Try Tp = 2700 K to check for energy balance:

Constituent CO, H2O N2

u2700 121 807.5 95 927.5 67 877.4 n 1 2 7.52 nu 121 807.5 191 854.9 510 438.1

Up(Tp) = 823 993.7 kJ

Page 216: Advanced Thermodynamics for Engineers

200 Thermodynamics of combustion

This is less than the value obtained from the energy equation and hence Tp > 2700 K. Try Tp = 2800 K:

U28W 127 154.9 100 470.8 70 754.1 n 1 2 7.52 nu 127 154.9 200 941.6 532 071.1

Up(Tp) = 860 167.6 kJ.

Hence, by linear extrapolation, try

887 571.8 - 860 167.6

860 167.6 - 823 993.7 T , = 2800 + x 100

= 2875.8 K

Constituent co2 H2O NZ

'2876 131 227.9 103 929.1 72 935 n 1 2 7.52 nu 131 227.9 207 858.1 548 471.5

U,(T,) = 887 557.5 kJ

This value is within 0.0016% of the value from the energy equation.

Example 4: combustion with a weak mixture of propane

A mixture of propane and excess air is burned at constant volume in an engine with a compression ratio of 10: 1. The strength of the mixture (q5) is 0.85 and the initial conditions at the start of the compression stroke are 1.2 bar and 127°C. Assume the ratio of specific heats for compression is 1.4 and take the internal energy of reaction ( Q , ) , of propane as -46 440 kJ/kg. Evaluate the final temperature and pressure after combustion neglecting dissociation. The internal energy of propane is 12 911.7 kJ/kmol at the beginning of compression and 106 690.3 kJ/kmol at the end (before combustion).

Solution

The stoichiomehic chemical equation is

C,H, + 5(02 + 3.76NZ)-3CO + 4H20 + l8.8NZ

The weak mixture equation is

5 C,H, + - (0, + 3.76NZ) + 302 + 4H2O + 0.88202 + 22.12N2

n oc

(10.33)

(10.34) np = 30.0 U . O J

nR = 29.0

Page 217: Advanced Thermodynamics for Engineers

Examples 201

The combustion equation (eqn (10.30)) is

= 1.2 x = 30.142 bar

K - 1

T , = Tl( :) = 400 x = 1004.75 K

The energy of reactants [U,(T,) - U,(T,)] is

Constituent C3H8 0 2 NZ

u l a ) 4 7 106 690.3 23 165.0 22 013.6 u298 12 91 1.7 6032.0 6025.8 Difference 93 778.7 17 133.0 15 987.9 n 1 5.882 22.12 nu 93 778.7 100 776.1 353 651.4

[UR(TR) - UR(T,)] = 548 206.2 W

The energy of products at standard temperature, T,, is

'298 7041.6 7223.3 6031.7 6025.8 n 3 4 0.882 22.12 nu 21 124.8 28 893.2 5320.0 133 290.7

Up(Ts) = 188 628.7 W

Up(Tp) = 46 440 x (12 x 3 + 8) + 548 206.2 + 188 628.7

= 2 043 360.0 + 548 206.2 + 188 628.7

= 2 780 194.9 kJ

Assume Tp = 3100 K:

u3100 143 420.0 114 194.9 84 974.72 79 370.95 n 3 4 0.882 22.12 nu 430 260.1 456 779.6 74 947.7 1 755 685.0

Up(Tp) = 2 717 672.4 W

Page 218: Advanced Thermodynamics for Engineers

202 Thermodynamics of combustion

Try Tp = 3200 K:

Constituent co2 H2 0 2 NZ

113200 148 954.8 118 773.6 88 164.1 82 226.2 n 3 4 0.882 22.12 nu 446 864.5 475 094.3 77 760.8 1 818 843

Up(Tp) = 2 818 562.6 W

Interpolating linearly gives

T , = 3100 + 2 780 194.9 - 2 717 672.4

2 818 562.6 - 2 717 672.4 x 100 = 3162 K

Check solution using Tp = 3 162 K:

Constituent co2 H2O 0 2 NZ

'3162 146 843.0 117 034.7 86 949.1 81 142.8 n 3 4 0.882 22.12 nu 440 529 468 138.7 76 689.4 1 794 878

Up(Tp) = 2 780 234.8 W.

This value is within 0.0014% of the value calculated from the energy equation. Hence Tp = 3162 K.

Pressure at the end of combustion is

p , = - x - 30 3162 x l o x 12 nPsTP Vl p 3 = - -

n ~ s T i V3 = 98.13 bar

29 400

Example 5: combustion at constant pressure of a rich mixture of benzene

A rich mixture of benzene (&I-&) and air with an equivalence ratio, @, of 1.2 is burned at constant pressure in an engine with a compression ratio of 15 : 1. If the initial conditions are 1 bar and 25°C calculate the pressure, temperature and volume after combustion. The ratio of specific heats during compression, K, is 1.4 and the calorific value of benzene is 40 635 kJ/kg. See Fig 10.9.

p 1 = 1 bar, Tl = 298 K 1.4

p 2 = p l ( z ) = 44.31 bar

T, = TI( zr = 880 K

Page 219: Advanced Thermodynamics for Engineers

Examples 203

Volume

Fig. 10.9 p - V diagram for constant pressure combustion

The stoichiometric chemical equation is

C,& + 7.5(O, + 3.76N,)46CO, + 3H,O + 23.5N2

The rich mixture equation is

7.5 1.2

C6H6 + - (0, + 3.76N2) - aCO, + bCO + 3H20 + 23.5N2

a = 3.5; b = 2.5 a + b = 6 h + b + 3 = 125

Hence

C6H6 + 6.25(0, + 3.76N2) - 3.5C02 + 2 3 2 0 + 3H,O + 23.5N, nR = 30.75 np = 32.5

The energy equation for constant pressure combustion is

Constituent C6H6 0 2 N2

h**o 39 936.1 27 230.9 26 342.1 h298 8201.8 8509.7 8503.4 Difference 31 734.4 18 721.2 17 838.7 n 1 6.25 23.5 nh 31 734.4 117 007.5 419 209.3

(10.35)

(10.36)

(10.37)

(10.38)

[HR(TR) - HR(T,)] = 567 951.2

Page 220: Advanced Thermodynamics for Engineers

204 Thermodynamics of combustion

The products enthalpy at standard temperature, T,:

Constituent COZ co H2O N2

h298 9519.3 8489.5 970 1 .O 8503.4 n 3.5 2.5 3 23.5 nh 33 317.5 21 223.8 29 102.9 199 830.6

Hp(T,) = 283 474.8 kJ

In this case the combustion is not complete and hence the full energy available in the fuel is not released. Applying Hess’ law, the energy released by a combustion process is given by

In this case

= 3.5(Ahf)COz + 2 . 5 W f I c 0 + 3(Ahf)H20 - = 3.5 x -393 522 + 2.5 x - 110 529 + 3 x -241 827 - (82 847.6) = -2 461 978.1 kJ (10.39)

This value can also be calculated from the enthalpy of reaction by subtracting the energy which is not released because of the incomplete combustion of all the carbon to carbon dioxide,

( Q p ) 2 5 = (Qp)25,total- 2.5( (Ahf)CO,- (Ahf)COl = -40635 x 78 - 2.5( -393 522 + 110 529)

= -3 169 530 + 707 482.5

= -2 462 047.5 kJ

These two values of enthalpy of combustion are the same within the accuracy of the data used (they differ by less than 0.003%). Hence, referring to Fig 10.7, the loss of energy available due to incomplete combustion is 707 482.5 kJ/kmol C,H, because not all the carbon is converted to carbon dioxide. Thus

Hp(Tp) = 2 462 047.5 + 567 888.1 + 283 471.3 = 3 313 406.9 W

Assume Tp = 2800 K:

Constituent COZ co H2O N2

~*, 150 434.9 94 792.4 123 750.8 94 034.2 n 3.5 2.5 3 23.5 nh 526 522.2 236 981.0 371 252.5 2 209 803

Hp(Tp) = 3 344 558.7 kJ

Page 221: Advanced Thermodynamics for Engineers

Problems 205

Hence Tp < 2800 K. Use Tp = 2700 K:

Constituent CO, co HZO NZ ~~~ ~ ~

h*7, 144 256.1 91 070.2 118 376.1 90 326.0 n 3.5 2.5 3 23.5 nh 504 896.3 227 675.4 355 128.2 2 122 661

Hp(Tp) = 3 210 360.9 kJ

Interpolate linearly:

3 313 406.9 - 3 210 360.9

3 344 558.7 - 3 210 360.9 Tp = 2700 + x 100 = 2776.8 K

Constituent CO, co H2O NZ

hZ777 149 011.4 93 936.1 122 512.5 93 181.1 n 3.5 2.5 3 23.5 nh 521 540 234 840.3 367 537.5 2 189 757

Hp(Tp) = 3 313 674.8 kJ

Hence the equations balance within 0.010%.

p3 = 44.31 bar. Hence the volume at the end of combustion, V,, is given by The pressure at the end of combustion is the same as that at the end of compression, i.e.

32.5 x 2777 1 v, = ___ x - x V , = nPsTP P1 x- x VI nRST1 p3 30.75 x 298 44.31

= 0.2223V1

10.7 Concluding remarks

A consistent method of analysing combustion has been introduced. This is suitable for use with all the phenomena encountered in combustion, including weak and rich mixtures, incomplete combustion, heat and work transfer, dissociation and rate kinetics. The method which is soundly based on the First Law of Thermodynamics ensures that energy is conserved.

A large number of examples of different combustion situations have been presented.

PROBLEMS

Assume that air consists of 79% N, and 21% 0, by volume.

1 Calculate the lower heat of reaction at constant volume for benzene (C,H,) at 25°C. The heats of formation at 25°C are: benzene (C,H,), 80.3 MJ/kmol; water vapour (H,O), -242 MJ/kmol; carbon dioxide (CO,), -394.0 MJ/kmol.

Page 222: Advanced Thermodynamics for Engineers

206 Thermodynamics of combustion

A mixture of one part by volume of vaporised benzene to 50 parts by volume of air is ignited in a cylinder and adiabatic combustion ensues at constant volume. If the initial pressure and temperature of the mixture are 10 bar and 500 K respectively, calculate the maximum pressure and temperature after combustion neglecting dissociation. (Assume the internal energy of the fuel is 8201 kJ/kmol at 298 K and 12 998 kJ/kmol at 500 K.)

[-3170 MJ/kmol; 52.11 bar; 2580 K]

The heat of reaction of methane (CH,) is determined in a constant pressure calorimeter by burning the gas as a very weak mixture. The gas flow rate is 70 litre/h, and the mean gas temperature (inlet to outlet) is 25°C. The temperature rise of the cooling water is 1.8"C, with a water flow rate of 5 kg/min. Calculate the higher and lower heats of reaction at constant volume and constant pressure in kJ/kmol if the gas pressure at inlet is 1 bar.

[-800 049.5 kJ/kmol; -887 952 kJ/km01]

In an experiment to determine the calorific value of octane (C8HI8) with a bomb calorimeter, the mass of octane was 5.421 95 x loT4 kg, the water equivalent of the calorimeter including water 2.677 kg and the corrected temperature rise in the water jacket 2.333 K. Calculate the lower heat of reaction of octane, in kJ/kmol at 15°C.

If the initial pressure and temperature were 25 bar and 15°C respectively, and there was 400% excess oxygen, estimate the maximum pressure and temperature reached immediately after ignition assuming no heat losses to the water jacket during this time. No air was present in the calorimeter.

[-5 099 000 kJ/km01; 3135 K]

A vessel contains a mixture of ethylene (C2H,,) and twice as much air as that required for complete combustion. If the initial pressure and temperature are 5 bar and 440 K, calculate the adiabatic temperature rise and maximum pressure when the mixture is ignited.

If the products of combustion are cooled until the water vapour is just about to condense, calculate the final temperature, pressure and heat loss per kmol of original mixture.

The enthalpy of combustion of ethylene at the absolute zero is -1 325 671 kJ/ kmol, and the internal energy at 440 K is 16 529 kJ/kmol.

[1631 K; 23.53 bar; 65°C; 3.84 bar; 47 119 kJ/kmol]

A gas engine is operated on a stoichiometric mixture of methane (CK) and air. At the end of the compression stroke the pressure and temperature are 10 bar and 500 K respectively. If the combustion process is an adiabatic one at constant volume, calculate the maximum temperature and pressure reached.

[59.22 bar; 2960 K]

A gas injection system supplies a mixture of propane (C,H8) and air to a spark ignition engine, in the ratio of volumes of 1 : 30. The mixture is trapped at 1 bar and 300 K, the volumetric compression ratio is 12 : 1, and the index of compression K = 1.4. Calculate the equivalence ratio, the maximum pressure and temperature achieved during the cycle, and also the composition (by volume) of the dry exhaust gas.

[0.79334, 119.1 bar, 2883 K; 0.8463, 0.1071, 0.04651

Page 223: Advanced Thermodynamics for Engineers

Problems 207

7 A turbocharged, intercooled compression ignition engine is operated on octane (C8H18) and achieves constant pressure combustion. The volumetric compression ratio of the engine is 20 : 1, and the pressure and temperature at the start of compression are 1.5 bar and 350 K respectively. If the air-fuel ratio is 24: 1, calculate maximum temperature and pressure achieved in the cycle, and the indicated mean effective pressure (imep, pi) of the cycle in bar. Assume that the index of compression K~ = 1.4, while that of expansion, K~ = 1.35

[2495 K; 99.4 bar; 15.16 bar]

8 One method of reducing the maximum temperature in an engine is to run with a rich mixture. A spark ignition engine with a compression ratio of 10: 1, operating on the Otto cycle, runs on a rich mixture of octane (C,H,,) and air, with an equivalence ratio of 1.2. The trapped conditions are 1 bar and 300 K, and the index of compression is 1.4. Calculate how much lower the maximum temperature is under this condition than when the engine was operated stoichiomehically. What are the major disadvantages of operating in this mode?

[208"C]

9 A gas engine with a volumetric compression ratio of 10 : 1 is run on a weak mixture of methane (CH,) and air, with an equivalence ratio @ = 0.9. If the initial temperature and pressure at the commencement of compression are 60°C and 1 bar respectively, calculate the maximum temperature and pressure reached during combustion at constant volume if compression is isentropic and 10% of the heat released during the combustion period is lost by heat transfer.

Assume the ratio of specific heats, K, during the compression stroke is 1.4, and the heat of reaction at constant volume for methane at 25°C is -8.023 x 10' kJ/kmol C&.

[2817 K; 84.59 bar]

10 A jet engine bums a weak mixture (@ = 0.32) of octane (C,H,,) and air. The air enters the combustion chamber from the compressor at 10 bar and 500 K; assess if the temperature of the exhaust gas entering the turbine is below the limit of 1300 K. Assume that the combustion process is adiabatic and that dissociation can be neglected. The enthalpy of reaction of octane at 25°C is -44 880 kJ/kg, and the enthalpy of the fuel in the reactants may be assumed to be negligible.

[Maximum temperature, Tp = 1298 K; value is very close to limit]

11 A gas engine is run on a chemically correct mixture of methane (CH,) and air. The compression ratio of the engine is 10 : 1, and the trapped pressure and temperature at inlet valve closure are 60°C and 1 bar respectively. Calculate the maximum temperature and pressure achieved during the cycle if:

(a) combustion occurs at constant volume; (b) 10% of the energy added by the fuel is lost through heat transfer; (c) the compression process is isentropic.

energy of methane in the reactants is negligible. It can be assumed that the ratio of specific heats K = 1.4, and that the internal

[2956 K; 88.77 bar]

Page 224: Advanced Thermodynamics for Engineers

11 Chemistry of Combustion

The thermodynamics of combustion were considered in Chapter 10, and it was stated that adiabatic combustion could be achieved. The concept of adiabatic combustion runs counter to the experience of many engineers, who tend to relate combustion to heat addition or heat release processes. This approach is encouraged in mechanical engineers by the application of the air standard cycle to engines to enable them to be treated as heat engines. In reality combustion is not a process of energy transfer but one of energy transformation. The energy released by combustion in a spark ignition (petrol) engine is all contained in the mixture prior to combustion, and it is released by the spark. It will be shown that the energy which causes the temperature rise in a combustion process is obtained by breaking the bonds which hold the fuel atoms together.

11.1

Heats (enthalpies and internal energies) of formation can be evaluated empirically by ‘burning’ the fuel. They can also be evaluated by consideration of the chemical structure of the compound. Each compound consists of a number of elements held together by certain types of bond. The bond energy is the amount of energy required to separate a molecule into atoms; the energy of a particular type of bond is similar irrespective of the actual structure of the molecule.

This concept was introduced in Chapter 10, and heats of formation were used to evaluate heats of reaction (Hess’ law). The process of breaking the chemical bonds during the combustion process can be depicted by a diagram such as Fig 1 1.1. It is assumed that element molecules can be atomised (in a constant pressure process) by the addition of energy equal to AHa. If these atoms are then brought together they would combine, releasing dissociation energy of c A H ( X - nR, to form the reactants. The sum of the dissociation and atomisation energies (taking account of the signs) results in the enthalpy of formation of the reactants. In a similar way the enthalpy of formation of the products can be evaluated. Using Hess’ law, the enthalpy of reaction of the fuel can be evaluated as the difference between the enthalpies of formation of the products and the reactants These energies are essentially the bond energies of the various molecules, and some of these energies are listed in Table 1 1.1.

Figure 11.2 shows how the energy required to separate two atoms varies with distance: the bond energy is defined as the minimum potential energy relative to that at infinity. The point of minimum energy indicates that the molecule is in equilibrium.

Bond energies and heats of formation

Page 225: Advanced Thermodynamics for Engineers

Gaseous atoms

Fig. 11.1 Relationship between enthalpies of formation and reaction

n n 4

energy

c m

Elemental n molecules

-

Distance

Fig. 11.2 Variation of bond energy with distance between atoms

CAH(X-Y), CAH(X-Y) ,

n n

m3, W f h

Table 11.1 Some atomisation, dissociation and resonance energies (based on 25OC)*

Reactant molecules

Bond Bond Energy dissociation Energy Energy atomisation (AH,) (MJ/kmol) ( A H ( X - Y ) ) (MJ/kmol) Resonance (MJ/kmol)

. V V V

H-H 435.4 C-H C (graphite) 717.2 N--H (O=O)o, 498.2 0-H N=N 946.2 H-OH

c-0 c=o c-c c=c c=c

414.5 Benzene: C,H, 150.4 359.5 Naphthalene: C,J& 255.4 428.7 Carbon dioxide: CO, 137.9 497.5 -COOH group 117.0 35 1.7 698.1 347.5 615.5 812.2

*note: these values have been taken from different sources and may not be exactly compatible.

Page 226: Advanced Thermodynamics for Engineers

2 10 Chemistry of combustion

In addition to the energy associated with particular bonds there are also energies caused by the possibility of resonance or strain in a particular molecule. For example, the ring structure of benzene, which is normally considered to be three simple double bonds between carbon atoms, together with three single ones, plus six carbon-hydrogen bonds, can reform to give bonding across the cyclic structure (see Fig 11.3).

This results in a substantially higher energy than would be estimated from the simple bond structure, and some typical values are shown in Table 1 1.1.

H H H H H

Fig. 11.3 Bonding arrangements of benzene, showing bond resonance through delocalised electrons

11.2 Energy of formation

It will be considered in this section that the processes take place at constant pressure and the property enthalpy ( h or H ) will be used; if the processes were constant volume ones then internal energies ( u or U) would be the appropriate properties.

A hydrocarbon fuel consists of carbon and hydrogen (and possibly another element) atoms held together by chemical bonds. These atoms can be considered to have been brought together by heating carbon and hydrogen (in molecular form) under conditions that encourage the resulting atoms to bond. There are two processes involved in this: first, the carbon has to be changed from solid graphite into gaseous carbon atoms and the hydrogen also has to be atomised; then, second, the atoms must be cooled to form the hydrocarbon fuel. These two processes have two energies associated with them: the first is atomisation energy (AH,) and the second is dissociation energy (X A H ( X - Y)R) required to dissociate the chemical bond in the compound CJ,. Figure 11.1 shows these terms, and also indicates that the energy of formation of the fuel ( A H , ) is the net energy required for the process, i.e.

A H , = AH, - C A H ( X - Y)R (11.1)

In reality the enthalpy of formation is more complex than given above because energy can be stored in molecules in a number of ways, including resonance energies (in the benzene ring structure) and changes of phase (latent heats). A more general representation of the enthalpy of formation is

(11.2) A H , = C AH, - A H ( X - Y ) - C AHms - C AH,,,,

Example

Evaluate the enthalpy of formation of CO, and H,O from the atomisation and dissociation energies listed in Table 1 1.1.

Page 227: Advanced Thermodynamics for Engineers

Energy of formation 21 1

Solution

Carbon dioxide (CO,) Carbon dioxide is formed from carbon and oxygen in the reaction.

cmli, + O,(g)-+CO,(g) (11.3)

where the (g) indicates that the element or compound is in the gaseous (vapour) phase, and (1) will be used to indicate that the element or compound is in the liquid phase.

The reaction in eqn (1 1.3) is achieved by atomisation of the individual carbon (graphite) molecules and the oxygen molecules, with subsequent recombination to form carbon dioxide. Effectively the reactant molecules, which are in a metastable state, are activated above a certain energy to produce atoms which will then combine to form the stable CO, molecule (see Fig 1 1.4).

reactants

stable

I

Progress of reaction

Fig. 11.4 Gibbs energy variation during a reaction

Hence, from Fig 11.1,

(AHf)coz = C AHa -C AH(X - Y ) - C AHRs - C AHIaent (11.4)

In this case there is a resonance energy but the latent energy (i.e. latent heat) is zero. Then

(AfZf)coz = C AHa[Cgraphie1 + AHa[O = 01 - 2(C = 0) - AH,,[CO,] = 717.2 + 435.4 - 2 x 698.1 - 137.9 (11.5) = -381.5 MJ/km01

The tabulated value is -393.5 MJ/kmoi.

Water (H,O) This is formed from the hydrogen and oxygen molecules in the reaction

+ ;O,(g)--H,o(g) Thus

(AHf)Hzo = C AHa -C AH(X - Y ) - C AH,, - C AH,,,,

(11.6)

(11.7)

Page 228: Advanced Thermodynamics for Engineers

212 Chemistry of combustion

In this case AHEs = 0 but AH,,,,, depends upon the phase of the water. First, consider the water is in vapour phase, when AH,,,,, = 0. Then

(AHf)H20 = X AH, - C AH(X - Y ) =AH,[H - HI + iX AH,[O = 01 - [H - OH] - [0 - HI (11.8) = 435.4 + $ (498.2) - 497.5 - 428.7 = -241.7 MJ/kmol

This compares well with the tabulated values of -241.6 MJ/kmol. If the water was in liquid phase then eqn (1 1.6) becomes

H*(g) +$02(g)-HH,0(1)

(AHf)H20(1) = (AHf)HiO(g) + (mw)HzOhfg = -241.7 - 18 x 2441.8/1000 = -285.65 MJ/kmol

The tabulated value is -285.6 MJ/kmol.

(11.9)

(11.10)

Example

Evaluate the enthalpy of formation of methane, C&.

Solution

The structure of methane is tetrahedral, and of the form shown in Fig 11.5 (a). Methane is often depicted in planar form as shown in Fig 1 1.5 (b).

H H

H- C - H

H

(4 (b)

Fig. 11.5 Atomic arrangement of methane molecule

Methane can be formed by the following reaction:

Cgraphire + 2H2(g)-C%(g) (11.11)

Page 229: Advanced Thermodynamics for Engineers

Energy of formation 213

Hence

(AHf)Ch = C AH, - C AH(X - Y ) - C AH,, - C AHI,,,, = AH,[C,,,,] + 2 AHa[H - HI - 4[H - C] = 717.2 + 2 x 425.4 -4 x 414.5 = -70 MJ/kmol

(11.12)

The tabulated value is -74.78 MJ/kmol, and the difference occurs because the energy of the bonds in methane is affected by the structure of the methane molecule, which results in attraction forces between molecules.

Example

Evaluate the enthalpy of formation of ethanol, C,H,OH.

Solution

The planar representation of the structure of ethanol is shown in Fig 1 1.6.

H H

I I I I

H - C - C - 0 - H

H H Fig. 11.6 Structure of ethanol molecule

Ethanol can be considered to be made up of the following processes:

This gives the following equation:

-2C(g) + 2 x 717.2, + 6H + 3 x 435.4 + 0 + 0.5 x 498 2 atomlsation energy fur 0

atomisaoun - energy fur C energy for H

- 5[C-H] - [C-C] - [C-01 - [O-HI -C,H,OH(g) + 2 x 717.2 + 3 x 435.4 + 0.5 x 498.2

- 5 x 414.5 - 347.5 - 351.7 - 428.7 -C*H,OH(g) - 210.7 MJ/kmol

(11.13)

Page 230: Advanced Thermodynamics for Engineers

214 Chemistry of combustion

Hence the enthalpy of formation of gaseous ethanol is -210.7 MJ/kmol. This is equivalent to -5.853 MJ/kg of ethanol. The value obtained from tables is about -222.8 MJ/kmol, and the difference is attributable to the slight variations in bond energy that occur due to the three-dimensional nature of the chemical structure.

Example

Evaluate the enthalpy ofreaction of methane.

Solution

This can be obtained either by using atomisation and dissociation energies, in a similar manner to that used to find the enthalpies of formation of compounds, or from the enthalpies of formation of the compounds in the reactants and products. Both methods will be used in this case. The reaction describing the combustion of methane is

(11.14)

The chemical structure of methane was given above. Hence the enthalpy of formation of the reactants

CH4(g) + 20z(g)-~Oz(g) + 2HzO(g)

(AHf)R = C AHa - C AH(X - Y ) - C AHms - C AH,,,,, = AH,[C,,,i,] + 2 AHa[H - HI + 2 AHa[O = 01 - 4[H - C] - 2[0 = 01

(1 1.15)

Note that the atomisation and dissociation energies of oxygen are equal (2 AHa[O = 01 = 2[0 = 01) and cancel out, i.e. the enthalpy of formation of oxygen is zero. This value of zero is assumed as a base level for all elements. Thus

(11.16) (AHf)R = (AHf)CH4 = 70 MJ/kmOl

Similarly for the products

The heat of reaction is given by

AHR = (AHf)P - (AHf)R = -381.7 - 2 x 241.7 - (-70) = -794.9 MJ/kmol

( 1 1.17)

(11.18)

This is close to the value of -802.9 MJ/kmol quoted as the lower enthalpy of reaction of methane in Table 11.2. If the higher heat of reaction of methane is required then eqn (1 1.14) becomes

(11.19) CH4W + 2o~(g)--coz(g) + 2Hz0(1)

Page 231: Advanced Thermodynamics for Engineers

Energy of formation 215

Example

Evaluate the lower enthalpy of reaction of benzoic acid (C,H&OOH). The planar diagram of its structure is shown in Fig 1 1.7.

0

C- 0- H

HC

HC (I f CH

H

Fig. 11.7 Structure of benzoic acid molecule

Solution

The reaction of benzoic acid with oxygen is

C,H,COOH(g) + 7.5O2(g)-7C0,(g) + 3H20(g) (1 1.20)

The easiest way to evaluate the enthalpy of reaction is from eqn (1 1.18):

AHR = (AHf)P - (AHf)R

The values of Hf were calculated above for CO, and H,O, and hence the only unknown quantity in eqn (1 1.18) is the enthalpy of formation of the benzoic acid:

(11.21) (AHf)C6H5CmH = C AH, - C AH(X - Y ) - C AH,, - C AH,,,,

The acid is in gaseous form before the reaction, see eqn (1 1.20), and thus AH,,,,, = 0. Substituting values into eqn (1 1.21) gives

(AHf)C6H5COOH = C AH, - C AH(X - Y ) - C AH,, = 7 AH,[C,,,,ite] + 3 AH,[H - HI + AH,[O = 01

-5[H - C] -4[C - C] - 3[C=C] - [ C = O ] - LC - O1 - Lo - - [AHreslC6& - [AHmslCOOH (1 1.22)

Hence

(AHf)C6HsCOOH = 7 x 717.2 + 3 x 435.4 + 498.2 - 5 x 414.5 - 4 x 347.5 - 3 x 615.5 - 698.1 - 351.7 - 428.7 - 150.4 - 117.0

= - 230.1 MJ /km~l (1 1.23)

Page 232: Advanced Thermodynamics for Engineers

216 Chemistry of combustion

The enthalpy of formation of the products is

(AHf)k' = (AHf)COz + (AHf)HzO

= 7 x ( -381 .5 )+3~ (-241.7) = -3395.5 MJ/kmol

Thus the enthalpy of reaction of benzoic acid is

(AHR)c ,H,c ,H = -3395.6 - (-230.1) = 3165.5 MJ/kmol

The tabulated value is 3223.2 MJ/kmol, giving an inaccuracy of about 2%.

(11.24)

(1 1.25)

11.3 Enthalpy of reaction

The enthalpies of reaction of some commonly encountered fuels are given in Table 11.2. These have been taken from a number of sources and converted to units consistent with this text where necessary. There are a number of interesting observations that can be made from this table:

0 the enthalpies of reaction of many of the hydrocarbon fuels on a basis of mass are very similar, at around 44000 kJ/kg;

0 the stoichiometric air-fuel ratios of many basic hydrocarbon fuels lie in the range

0 some of the fuels have positive enthalpies of formation; 0 all of the fuels have negative enthalpies of reaction; 0 the enthalpies of reaction of the alcohols are less than those of the non-oxygenated

fuels, simply because the oxygen cannot provide any energy of reaction; 0 the commonly used hydrocarbon fuels are usually mixtures of hydrocarbon com-

pounds.

13:1-17:1;

11.4 Concluding remarks

It has been shown that the energy released by a fuel is contained in it by virtue of its structure, i.e. the bonds between the atoms. It is possible to assess the enthalpies of formation or reaction of a wide range of fuels by considering the chemical structure and the bonds in the compound.

A table of enthalpies of formation and reaction for common fuels has been given.

Page 233: Advanced Thermodynamics for Engineers

mw

bN

mm

dm

d

I

N

d-

ow

m wom

ma

00

o

mm

m

w+

w

0-

wo

m

me

m

mm

m~

r

-2

~

Il

l1

I

I

Il

l

NN

mb

+

wr

-

wm

mm

O

D0

m

oo

m

mb

-

No

m0

+

mm

Il

l

wm

*z

o

mm

C

IN

N

++

++

I

ll

1

s

Page 234: Advanced Thermodynamics for Engineers

12 Chemical Equilibrium and Dissociation

Up to now this book has concentrated on combustion problems which can be solved by methods based on equilibrium but which do not require an explicit statement of the fact, e.g. complete combustion of a hydrocarbon fuel in air can be analysed by assuming that the products consist only of H,O and CO,. These methods are not completely correct and a more rigorous analysis is necessary to obtain greater accuracy.

Consider the combustion of carbon monoxide (CO) with oxygen (02); up until now the reaction has been described by the equation

co + 1/20,-+CO, (12.1)

It is implied in this equation that carbon monoxide combines with oxygen to form carbon dioxide, and as soon as that has happened the reaction ceases. This is not a true description of what happens in practice. The real process is one of dynamic equilibrium with some of the carbon dioxide breaking down into carbon monoxide and oxygen (or even more esoteric components) again, which might then recombine to form carbon dioxide. The breakdown of the CO, molecule is known as dissociation. To evaluate the amount of dissociation that occurs (the degree of dissociation) it is necessary to evolve new techniques.

12.1 Gibbs energy

The concept of Gibbs energy, G, was introduced im Chapter 1. The change in the specific Gibbs energy, g, for a system of fixed composition was defined in terms of other properties as

(12.2)

It was also shown that for a closed system at constant temperature and pressure,

dg= v dp - s dT

performing only mechanical work, to be in equilibrium,

dG),,, = 0 (12.3)

Equations (12.2) and (12.3) are based on the assumption that G = mg = mg( p , T ) , and this is quite acceptable for a single component system, or one of fixed composition. If the system has more than one component, and these components can react to form other

Page 235: Advanced Thermodynamics for Engineers

Gibbs energy 219

compounds, e.g. if the system contained carbon monoxide, oxygen and carbon dioxide as defined in eqn (12.1), then it is necessary to define the Gibbs energy as G=mg=mg(p , T,mi) where m, is the mass of component i, and m = C mi. The significance of changes of composition on the value of the Gibbs energy of a mixture will now be investigated.

If

G = mg = mg( p, T, m,) (12.4)

and if it is assumed that G is a continuous function with respect to p, and T and the masses of constituents comprising the mixture, then the change of G with changes in the independent variables is

dG = ( $)T,dp + ( $)p,ciT + (”) dm, + ... (*) drn, (12.5a) a m l p,T,m, , l amn p.T.m,,,

where dm, . . . dm, are changes in mass of the various constituents. A similar equation can be written in terms of amount of substance n, and is

For the initial part of the development of these equations the mass-based relationship (eqn (12.5a)) will be used because the arguments are slightly easier to understand. The term

represents the ‘quantity’ of Gibbs energy introduced by the transfer of mass dm, of constituent 1 to the system. (This can be more readily understood by considering the change in internal energy, dU, when the term (aU/am,)p,T,m,*, dm, has a more readily appreciated significance.)

The significance of the terms on the right of eqn (12.5a) is as follows:

1. The first term denotes the change of Gibbs energy due to a change in pressure, the temperature, total mass and composition of the system remaining constant. This is equivalent to the term derived when considering a system of constant composition, and is V dp. The second term denotes the change of Gibbs energy due to a change in temperature, the pressure and total mass of the system remaining constant. This is equivalent to - S dT derived previously. The third term shows the change of Gibbs energy due to a change in the mass (or amount of substance if written in terms of n) of constituent m,, the pressure, temperature and masses of other constituents remaining constant. It is convenient to define this as

2.

3.

PI =(E) 4. The fourth term is a general term of the form of term (3) in eqn (12.6).

(12.6)

Page 236: Advanced Thermodynamics for Engineers

220 Chemical equilibrium and dissociation

Hence, in terms of masses

d G = V d p - S d T + x p i d m i i = 1

while in terms of amount of substance n

dG = Vdp - SdT + ~ p , , d n ,

(12.7a)

(12.7b)

12.2 Chemical potential, p The term p is called the chemical potential and is defined as

( E ) p , T . m , * I

The significance of p will now be examined. First, it can be considered in terms of the other derived properties.

By definition

d G = d ( H - TS)=dH - T dS - S dT

Hence d H = dG + T dS+S dT

= V dp - S dT+ C p i dmi + T d S + S dT = V dp + T d S + C p i dmi

(12.8)

(12.9)

Considering each of the terms in eqn (12.9), these can be interpreted as the capacity to do work brought about by a change in a particular property. The first term is the increase in capacity to do work that is achieved by an isentropic pressure rise (cf. the work done in a feed pump of a Rankine cycle), and the second term is the increased capacity to do work that occurs as a result of reversible heat transfer. The third term is also an increase in the capacity of the system to do work, but this time it is brought about by the addition of a particular component to a mixture. For example, if oxygen is added to a mixture of carbon monoxide, carbon dioxide, water and nitrogen (the products of combustion of a hydrocarbon fuel) then the mixture could further react to convert more of the carbon monoxide to carbon dioxide, and more work output could be obtained. Thus p i is the increase in the capacity of a system to do work when unit mass (or, in the case of pmi, unit amount of substance) of component i is added to the system. p i can be considered to be a ‘chemical pressure’ because it is the driving force in bringing about reactions.

Assuming that H is a continuous function,

(12.10) By comparison of eqns (12.9) and (12.10)

(12.11)

Page 237: Advanced Thermodynamics for Engineers

Stoichiometry 221

Similarly it can be shown that

The following characteristics of chemical potential may be noted:

( 1 2.1 2)

(i)

(ii)

the chemical potential, p, is a function of properties and hence is itself a thermodynamic property; the numerical value of p is not dependent on the property from which it is derived (all the properties have the dimensions of energy and, hence, by the conservation of energy this is reasonable);

(iii) the numerical value of p is independent of the size of the system and is hence Gn intensive property, e.g.

(12.13)

Since p is an intensive property it may be compared with the other intensive properties p and T etc. By the two-property rule this means that

P = P(P9 T ) (12.14)

and similarly

P = p ( PP T ) (12.15)

It can be shown that the chemical potential p for a pure phase is equal in magnitude to the specific Gibbs energy at any given temperature and pressure, i.e.

l u = g (12.16)

[Note: Although p =g, it is different from g inasmuch as it is an intensive property whereas g is a specific property. Suppose there is a system of mass m, then the total Gibbs energy is G = mg whereas the chemical potential of the whole system is still p { cf. P or 7-1.1

12.3 Stoichiometry

Consider the reaction

1

2 co + -02 * co, (12.17)

Equation (12.17) shows the stoichiometric proportions of the reactants and products. It shows that 1 mol CO and $mol 0, could combine to form 1 mol CO,. If the reaction proceeded to completion, no CO or 0, would be left at the final condition.

This is the stoichiometric equation of the reaction and the amounts of substance in the equation give the stoichiometric coefficients.

Page 238: Advanced Thermodynamics for Engineers

222 Chemical equilibrium and dissociation

The general equation for a chemical reaction is

v ,A + v b B e v,C + vdD (12.18)

where v is a stoichiometric coefficient, and A, B, C and D are the substances involved in the reaction. Applying eqn (12.18) to the CO + $0, reaction gives

vco = - 1 vq = -112 vcq= 1

(12.19)

It is conventional in chemistry to assign negative values to the stoichiometric coefficients on the left-hand side of the equation (nominally, the reactants) and positive signs to those on the right-hand side (nominally, the products).

12.3.1 MIXTURES

Mixtures are not necessarily stoichiometric and the following terms were introduced in Chapter 10 to describe the proportions of a mixture:

(i) If reactants occur in proportion to the stoichiometric coefficients then the mixture is said to be chemically correct or stoichiometric.

(ii) If the reactants have a greater proportion of fuel than the correct mixture then it is said to be rich.

(iii) If the reactants have a lesser proportion of fuel than the correct mixture then it is said to be weak.

e.g. chemically correct mixture is CO + O2 * C 0 2

Rich mixture (excess of fuel)

1

2 (1 + n)CO + - O2 a C 0 2 + nCO

Weak mixture (excess of oxidant)

C O + - + n O2*CO2+nO2 (: 1

(12.20)

(12.21)

NI3. Equations (12.20) and (12.21) have been written neglecting dissociation.

12.4 Dissociation

The basis of dissociation is the atomic model that all mixtures of gases are in a state of dynamic equilibrium. Molecules of the compounds are being created whilst existing ones are breaking down into simpler compounds or elements (dissociating). In the equilibrium situation the rates of creation and destruction of molecules of any compounds are equal. This means that macroscopic measuring techniques do not sense the changes but give the impression that the system is in a state of ‘static’ equilibrium. The effect of this is that the

Page 239: Advanced Thermodynamics for Engineers

Dissociation 223

reactions can no longer be said to be unidirectional but must be shown as

(12.22) 1

2 co + - o,aco,

On a molecular basis the above reaction can go either way. It is now necessary to consider a technique which will define the net direction of change for a collection of molecules. First, consider the general equation for the CO + io, reaction, neglecting particularly esoteric and rare compounds:

(12.23)

where the carbon ( C ) and atomic oxygen (0) are formed by the breakdown of the reactants. Experience shows that in the ranges normally encountered in practice, the C and 0 atoms have a negligible effect. This allows the general reaction to be simplified to

1

2 CO + - O , a a C O , + bCO + do, + eC +fO

1

2 co + - 0, a a C O , + bCO + do, (12.24)

It is possible to write this equation in a slightly different form by considering the amount of CO, that has dissociated. This can be defined as a = (1 - a ) , and the equation for the dissociation of CO, is

a 2

aCO,-aCO + - 0, (12.25)

By considering the stoichiometric equation (eqn (12.22)) and the dissociation equation (eqn (12.25)) a general equation may be constructed in terms of a:

co + 1/20,* co, aCO,-aCO + - 0,

2

Adding eqns (1 2.26) gives

(12.26)

1 a 2 2

CO + -0, 3 (1 - a)CO, + aCO+ - 0, (12.27)

In eqn (12.27), a is known as the degree ofdissociation. This equation shows the effect of dissociation on a chemically correct mixture. Before discussing methods of evaluating a, the effect of dissociation on non-stoichiometric mixtures will be shown, first using the carbon monoxide reaction, and then a general hydrocarbon fuel.

12.4.1 WEAK MIXTURE WITH DISSOCIATION Equation without dissociation:

CO+ - + n O , ~ C O , + n O , (: 1 (12.28)

(12.29)

Dissociation of CO,: a 2

a C 0 , e a C O + - 0,

Page 240: Advanced Thermodynamics for Engineers

224 Chemical equilibrium and dissociation

Adding eqns (12.28) and (12.29) gives

O , + ( l - a ) C O , + a C O +

12.4.2 RICH MIXTURE WITH DISSOCIATION

Equation without dissociation:

1

2 (1 + n)CO + - 0 2 =3 co, + nCO

Dissociation of CO,:

(12.30)

(12.31)

a 2

aCO2-aCO + - 0,

Total reaction:

1 a 2 2

(1 + n)CO + - 0, =3 (1 + a)CO, + (n + a)CO + - 0, (12.32)

12.4.3

A general hydrocarbon fuel can be defined as C,H, and this will react with the stoichiometric quantity of air as shown in the following equation:

GENERAL HYDROCARBON REACTION WITH DISSOCIATION

(12.33) ( 3 Y C, H , + x + - (0, + 3.76 N2) + xCO, + - H,O + 3.76 x + - N , ( :I 2

The dissociation of the CO, and H,O can be added into this equation as (eqn (12.29))

a1 2

a , C O , u a , C O + - 0,

and

a2 2

a ,H ,Oua ,H,+ - 0,

which gives a general equation with dissociation

C,H, + x + - (0, + 3.76N2) ( Y Y 2 2

+ x ( l - a,)CO, + - (1 - a,)H,O +xa,CO + - a,H,

+ - a l + - a 2 0 , + 3 . 7 6 x + - N, (1 : ) ( :)

(12.34)

(12.35)

Page 241: Advanced Thermodynamics for Engineers

Calculation of chemical equilibrium and the law of mass action 225

If the fuel was benzene (C,I-&) then eqn (12.35) would become

cfjI-& + 7.5(O2 + 3.76N2)

*6(1- a,)C02 + 3(1- a,)H,O + 6a,CO + 3a2H2+ (3a, + 1.5a2)02 + 28.2N2

(12.36)

If the mixture were not stoichiometric then eqn (12.33) would be modified to take account of the air-fuel ratio, and eqns (12.35) and (12.36) would also be modified. These equations are returned to in the later examples. In eqn (12.36) the combination of nitrogen and oxygen has been neglected. In many combustion processes the oxygen and nitrogen join together at high temperatures to form compounds of these elements: one of these compounds is nitric oxide (NO), and the equations can be extended to include this reaction. This will also be introduced later.

12.4.4 GENERAL OBSERVATION

As a result of dissociation there is always some oxidant in the products; hence dissociation always reduces the effect of the desired reaction, e.g. if the reaction is exothermic then dissociation reduces the energy released (see Fig 12.1).

/ Reactants

Fig. 12.1 Effect of dissociation on combustion

Having introduced the concept of dissociation it is necessary to evolve a method that allows the value of the degree of dissociation, a, to be calculated. This method will be developed in the following sections.

12.5 Calculation of chemical equilibrium and the law of mass action

General relationships will be derived, and the particular case of the CO + io2 reaction will be shown in brackets { ).

It was previously shown that for a system at constant pressure and temperature to be in an equilibrium state it must have a minimum value of Gibbs energy i.e. dC),,, = 0.

Page 242: Advanced Thermodynamics for Engineers

226 Chemical equilibrium and dissociation

But, by definition, for a system at constant pressure and temperature

dG),., = p i dm, + p2 dm, + " ' p , dm, (12.37)

where m, , m2 ... m, are the masses (or amounts) of the possible constituents of the mixture.

Only four constituents will be considered during this discussion, two reactants and two products, but the theory can be extended to any number of constituents. The equilibrium equation for the complete reaction is

t I C 0 + 1 0,-co2 1 (12.38)

At some intermediate stage in the reaction the state may be represented as

Y, A + Y b B * (1 - &)Y, A + (1 - &)YJ f &VcC f E V d D

(12.39)

where stoichiometric coefficients equal to unity are implicit. E is known as the fraction ofreaction and is an instantaneous value during the reaction

as opposed to a , the degree of dissociation, which is a final equilibrium value. The use of E allows the changes in Gibbs energy to be considered as the reaction progresses.

For the purposes of evaluating the dissociation phenomena it is possible to consider the reaction occurring at constant temperature and pressure (equal to values obtained by other means or calculated by an implicit iterative technique). The Gibbs energy of the system may be described by the following equation:

G = ( l - E) Yap, + ( l - E ) Y b p b + E v c p c + E V d p d ]

(12.40)

To find the equilibrium condition while maintaining pressure and temperature constant, this function has to be minimised with respect to E , i.e. it is necessary to locate when

Now

dG = V dp - S dT+ C p , dm, (12.41)

and

dG),,, = p l dm, + p2 dm, + ... p,, dm, (12.42)

From here on, the mass form of the equation (eqn 12.5a) will be replaced by the mole form of the equation (eqn 12.5b), because this is more appropriate for chemical reactions.

Page 243: Advanced Thermodynamics for Engineers

Calculation of chemical equilibrium and the law of mass action 227

It is possible to relate the amount of substance of each constituent in terms of E , namely

n,= (1 - E ) Y , + A

n, = EV, + C nb = (1 - &)vb + B

( 12.43) nd= &Vd+D

where A, B, C and D allow for the excess amount of substance in non-stoichiometric mixtures. Hence

dn, = - v, dE (12.44)

and applying similar techniques

( 12.45)

This is known as the equation of constraint because it states that the changes of amount of substance (or mass) must be related to the stoichiometric equation (i.e. changes are constrained by the stoichiometry). Hence

giving

= -vapa - v b p b + v c P c + v d p d

= 0 for equilibrium (1 2.47)

i.e. at equilibrium

+ v b p b = v c P c + v d p d (12.48)

Since = g it is possible to describe ,u in a similar way to g, namely

,u = po+ %T In pr (12.49)

where

Po = Po + PU-1 (12.50)

and is called the standard chemical potential, and is the value of p at temperature T and the datum pressure, po.

The value of pressure to be used in eqn (1 2.49) is the ratio of the partial pressure of the individual constituent to the datum pressure.

Substituting for p in the equilibrium equation, eqn (12.48)

va(p: + BT In p r o ) + v b ( P l + ST In prb) = vc(p:+ ST In p r c ) + vd(p:+ ST In p r d )

(12.51) which can be rearranged to give

( v a P ~ + v b p ~ - ~ ~ P ~ - - d P ~ ) + S ~ ( v ~ I n p r o + vblnpr,-vc ~ p r c - v d I n p r d ) = 0 (12.52)

Page 244: Advanced Thermodynamics for Engineers

228 Chemical equilibrium and dissociation

Hence

1

The left-hand side of eqn (12.53) is the difference in the standard chemical potentials at the reference pressure p0 of 1 bar or 1 atmosphere. This is defined as

(12.54) -AG: = yap: + vbp; - v,,u:- vdp:

Hence

giving

(12.55)

(12.56)

K p r is called the equilibrium consfanf, and is a dimensionless value. Sometimes the equilibrium constant is defined as

(12.57)

i.e.

Kpr = K,pgu + ' b - 'c " d ) or K p = Kpr/pgo + ' b " c - " d ) (12.58)

K , has the dimensions of pressure to the power of the sum of the stoichiometric coefficients, i.e. p". The non-dimensional equilibrium constant is defined as

AG; In K,,, = - 3iT

(1 2.59)

Now AG: is a function of T alone (having been defined at a standard pressure, p o ) , therefore K, = f ( T ) . Equation (12.56) shows that the numerical value of K , is related to the datum pressure used to define pO. If the amounts of substance of reactants and products

Page 245: Advanced Thermodynamics for Engineers

Variation of Gibbs energy with composition 229

are the same then the value of K, is not affected by the datum pressure (because va + vb - v , - vd = 0). However, if the amounts of substance of products and reactants are not equal, as is the case for the CO + :O,-CO, reaction, then the value of K, will be dependent on the units of pressure. Hence, the value of equilibrium constant, K,, is the same for the water gas reaction (CO+H,O-CO,+H,) in both SI and Imperial units because Kp is dimensionless in this case.

12.5.1

The definition of partial pressure is

K , DEFINED IN TERMS OF MOLE FRACTION

P a = X a P ( 1 2.60)

Hence, replacing the terms for partial pressure in eqn (12.57) by the definition in eqn (12.60) gives

(12.61)

which results in the following expression for K, in terms of mole fraction:

(12.62)

The above expressions are known as the law of mass action.

12.6 Variation of Gibbs energy with composition Equations (12.57) and (12.62) show that the equilibrium composition of a mixture is defined by the equilibrium constant, which can be defined in terms of the partial pressures or mole fractions of the constituents of the mixture; the equilibrium constant was evaluated by equating the change of Gibbs energy at constant pressure and temperature to zero, i.e. dG),,,=O. It is instructive to examine how the Gibbs energy of a mixture varies with composition at constant temperature and pressure. Assume that two components of a mixture, A and B , can combine chemically to produce compound C. This is a slightly simplified form of eqns (12.38) and (12.39). If the chemical equation is

A+B*2C (12.63)

then at some point in the reaction, defined by the fraction of reaction, E , the chemical composition is

( 1 - & ) A + ( 1 - E ) B + 2&C (12.64)

and the Gibbs energy is

G = ( 1 - E ) p a + ( 1 - & ) p b + 2 & p , ( 1 2.65)

This can be written, by substituting for p from eqns (12.49) and (12.50), as

G = ( 1 - E ) ~ : + % T l np ro+ ( 1 - &),DE +%T In prh+2&(p;+%T In p,,) = [ ( l - ~ ) p t + ( 1 - E),U;+ 2~p,0] + % T [ ( l - &)ln p,, + ( 1 - &)ln prb+ 2~ In p,,]

(12.66)

Page 246: Advanced Thermodynamics for Engineers

230 Chemical equilibrium ana' dissociation

The partial pressures are defined by eqn (12.60) as p i = x i p , and the mole fractions of the constituents are

I - & 1 - E ; x , = - = E 2E 2

xa = -; ' b = - 2 2

Substituting these terms in eqn (12.66) gives

= [(l - + (1 - &)p: + 2Ep3 + rnT[(l - E ) + (1 - E ) + 2 E I h p ,

(1 2.67)

(12.68)

Equation (12.68) can be rearranged to show the variation of the Gibbs energy of the mixture as the reaction progresses from the reactants A and B to the products C by subtracting 2p: from the left-hand side, giving

G - 2p: = (1 - &)[/A; +p; - 2p:] + 2RTlnpr + 2%T [ (1 - &)ln (';')+&h&] -

(12.69)

Equation (12.69) consists of three terms; the second one simply shows the effect of pressure and will be neglected in the following discussion. The first term is the difference between the standard chemical potentials of the separate components (p: + pt) before any reaction has occurred, and the standard chemical potential of the mixture (2~:) after the reaction is complete. Since the standard chemical potentials are constant throughout this isothermal process, this term varies linearly with the fraction of reaction, E. The third term defines the change in chemical potential due to mixing, and is a function of the way in which the entropy of the mixture (not the specific entropy) varies as the reaction progresses. The manner in which the first and third terms might vary is shown in Fig 12.2, and the sum of the terms is also shown. It can be seen that for this example the equilibrium composition is at E = 0.78. This figure illustrates that the Gibbs energy of the mixture initially reduces as the composition of the mixture goes from A + B to C. The standard chemical potential of compound C is less than the sum of the standard chemical potentials of A and B, and hence the reaction will tend to go in the direction shown in eqn (12.63). If the standard chemical potentials were the only parameters of importance in the reaction then the reactants A and B would be completely transformed to the product, C. However, as the reaction progresses, the term based on the mole fractions varies non-monotonically, as shown by the line labelled

2%+1 -&)h(+)+ &h&],

and this affects the composition of the mixture, which obeys the law of mass action. If the two terms are added together then the variation of G - 2pz with E is obtained. It can be

Page 247: Advanced Thermodynamics for Engineers

Examples of significance of K p 23 1

seen that the larger the difference between the standard chemical potentials (i.e. the steeper the slope of the line PQ) then the larger is the fraction of reaction to achieve an equilibrium composition. This is to be expected because the driving force for the reaction has been increased. Some examples of dissociation are given later, and these can be more readily understood if this section is borne in mind.

Standard Gibbs Energy of mixing of A and B

2RT[(l-&)ln((l-~)n)I~~]

Gibbs Energy ? of mixture

-2

0.00 0.50 1.00

Fraction of reaction

Fig. 12.2 Variation of the Gibbs energy of a mixture

12.7 Examples of significance of Kp

12.7.1 EXAMPLEI

Consider the reaction in eqn (12.1),

co+-oo,*co,

vco= -1, v q = -l /2,vco2=1

1

2

Hence

(12.70) Prco,

KPr = Prco &

Thus, if K, is known, the ratio of the partial pressures in the equilibrium state is known.

Page 248: Advanced Thermodynamics for Engineers

232 Chemical equilibrium and dissociation

It would be convenient to manipulate this expression into a more useful form. Consider the general reaction equation for the carbon monoxide and oxygen reaction (eqn (12.27)):

1 a 2 2

CO + - 0,-(1- a)CO, + a C O + - 0,

then

nco, p 1-a p =-- Prco, = - - nP Po 1 + a / 2 Po

(12.71)

Hence, the value of K , can be related to the degree of dissociation, a , through the following equation

(1 2.72a) l + a / 2 a

and

(12.72b)

The relationship between Kh and a allows the degree of dissociation, a , to be evaluated if K, is known. Equations (12.72) shows that, for this reaction, the value of K, is a function of the degree of dissociation and also the pressure of the mixture. This is because the amount of substance of products is not equal to the amount of substance of reactants.

12.7.2 EXAMPLE 2

Consider the water-gas reaction

CO,+H,eCO+H,O

By the law of mass action

(1 2.73)

(12.74)

In this reaction the amount of substance in the products is equal to the amount of substance in the reactants, and there is no effect of pressure in the dissociation equation. Comparing the results for the carbon monoxide, (eqn (12.1)) and the water gas reactions (eqn (12.73)), it can be seen that if a mixture of products of the reaction in eqn (12.1) was

Page 249: Advanced Thermodynamics for Engineers

Examples of significance of K, 233

subjected to a change in pressure, the chemical composition of the mixture would change, whereas the chemical composition of the products of the water gas reaction would be the same at any pressure.

These points are returned to later in this section.

12.7.3

A spark ignition engine operates on a 10% rich mixture of carbon monoxide (CO) and air. The conditions at the end of compression are 8.5 bar and 600 K, and it can be assumed that the combustion is adiabatic and at constant volume. Calculate the maximum pressure and temperature achieved if dissociation occurs.

[There are two different approaches for solving this type of problem; both of these will be outlined below. The first approach, which develops the chemical equations from the degrees of dissociation is often the easier method for hand calculations because it is usually possible to estimate the degree of dissociation with reasonable accuracy, and it can also be assumed that the degree of dissociation of the water vapour is less than that of the carbon dioxide. The second approach is more appropriate for computer programs because it enables a set of simultaneous (usually non-linear) equations to be defined. ]

EXAMPLE I : A ONE DEGREE OF DISSOCIATION EXAMPLE

General considerations

The products of combustion with dissociation have to obey all of the laws which define the conditions of the products of combustion without dissociation, described in Chapter 10, plus the ratios of constituents defined by the equilibrium constant, K,. This means that the problem becomes one with two iterative loops: it is necessary to evaluate the degree of dissociation from the chemical equation and the equilibrium constant, and to then ensure that this obeys the energy equation (i.e. the First Law of Thermodynamics). The following examples show how this can be achieved. At this stage the value of K , at different temperatures will simply be stated. The method of calculation K, is described in section 12.8, and K, values are listed in Table 12.3.

Solution

Note that there is only one reaction involved in this problem, and that the carbon monoxide converts to carbon dioxide in a single step. Most combustion processes have more than one chemical reaction.

Stoichiometric combustion equation:

1 CO + - (Q + 3.76N2) 3 CO, + 1.88 N2

n p = 2.88 L

n R = 3.38

Rich combustion equation:

1

2 1.1CO + - (Q + 3.76N2) C 0 2 + 0.1CO + l.88N2

3

np = 2.98 nR = 3.48

(12.75)

(12.76)

Page 250: Advanced Thermodynamics for Engineers

234 Chemical equilibrium ana‘ dissociation

Dissociation of C02 (eqn (12.29)):

a 2

a C 0 2 a a C 0 + - 0,

Total combustion equation with dissociation

1 a 2 2

l . l C O + - ( Q + 3 . 7 6 N 2 ) - ( 1 - a ) C O , + ( O . l + a ) C O + - 0 2 + 1.88N2

nR = 3.48 np = 2.98 + a12

(12.77)

The equilibrium constant, K,, is given by eqn (12.70):

(P/Po)co2

(P/Po)co <P/Po)bI: KP, =

The values of partial pressures in the products are

--xco2 - = - - 9

P2 0.1-a P2 - = xco - = - -. Pro2 P2 1 - a P 2 . PCO -- nP Po Po Po nP Po Po Po

Po, P2 a P2 =xo2 - = - -

Po Po 2nP Po - (12.78)

Also, from the ideal gas law,

p2V2 = n2%T2, and p lV l = nl%T, (12.79)

Thus

np n , T , 245.65 _ - _-=-

P2 PI 7-2 7-2

Hence, substituting these values in eqn (12.70) gives

( 1 -4 2nP Po (1 - a)’ 2 x 245.65 ( aT2 ”) (- -)’ a K:r = KPr = (0.1 + a ) a p 2 (0.1 + a)’

(12.80)

(12.81)

This is an implicit equation in the product’s temperature, T2, because K , = f (T2) . Writing Ki, as X gives

X(O.l + a)’aT2 = ( 1 - a)’x 2 x 2 4 5 . 6 5 ~ ~ = ( 1 - a)’ x 2 x 245.65 x 1.013 25 (12.82)

Expanding eqn (12.82) gives

0 = 497.77 - a(995.54 + 0.01XT2) + a’(497.77 - O.2XT2) - a3XT2 (12.83)

Page 251: Advanced Thermodynamics for Engineers

Examples of significance of K p 235

The term XT, can be evaluated for various temperatures because XT, = G T , :

2800 6.582 121 303 2900 4.392 55 940 3000 3.013 27 234

Solving for a at each temperature gives

TP a

2800 0.090625 2900 0.1218125 3000 0.171875

These values of a all obey the chemical, i.e. dissociation, equation but they do not all obey the energy equation. It is necessary to consider the energy equation now to check which value of T, balances an equation of the form

UP(',)= -(e,>,+ [ U R ( T R ) - ~ R ( T , ) I f UP(T,) This energy equation, based on the internal energy of reaction at T = 0, may be rewritten

0 = -Avo - Up(Tp) + U,(TR) (12.84) NOW UR(TR) is constant and is given by

Constituent co 0 2 NZ

~ s o o 12 626.2 12 939.6 12 571.7 n 1 .1 0.5 1.88

UR(TR) = 43 993.4 kJ

Hence,

Up(Tp) = (1 - a) x 276 960 + 43 993 kJ (12.85)

It can be seen that the value of Up is a function of a; the reason for this is because the combustion of the fuel (CO) is not complete when dissociation occurs. In a simple, single degree of freedom reaction like this, the reduction in energy released is directly related to the progress of the reaction:

TP a UPVP)

2800 0.090625 295 853 2900 0.1218125 287 216 3000 0.171875 273 351

Evaluating the energy which is contained in the products at 2900 K and 3000 K, allowing for the variation in a as the temperature changes, gives

Page 252: Advanced Thermodynamics for Engineers

236 Chemical equilibrium and dissociation

Tp = 2900 K

Constituent C 0 2 co 0 2 N*

u2900 131 933.3 74 388.9 78 706.6 73 597.5 n 0.871875 0.2281 0.0641 1.88

U,(Tp) = 275 405 kJ

Tp = 3000 K

Constituent C 0 2 CO 0 2 N2

u3wo 137 320 77 277 81 863 76 468 n 0.8281 0.2719 0.085 94 1.88

Lrp(Tp) = 285 521 kJ

These values are plotted in Fig 12.3, and it can be seen that, if the variation of the energy terms was linear with temperature, the temperature of the products after dissociation would be 2949 K. The calculation will be repeated to show how well this result satisfies both the energy and dissociation equations.

288000 , 286000

284000 az . * v) 282000

z = 280000 e, 9 '- 278000

8 276000 w

274000

LlLVYY ~ ~

2900 2910 2920 2930 2940 2950 2960 2970 2980 2990 3000

Temperature / (K)

Fig. 12.3 Energy contained in products based on eqn (12.80) and the tables of energies

First, it is necessary to evaluate the degree of dissociation that will occur at this product temperature. At Tp = 2950 K, Kpr = 3.62613 (see Table 12.3), and this can be substituted into eqn (12.78) to give a = 0.1494.

Hence, the chemical equation, taking account of dissociation, is

1

2 1.1CO + - ( 0 2 + 3.76 N2) 0.8506 CO, + 0.2494 CO + 0.0747 0 2 + 1.88 N2

Applying the energy equation: the energy released by the combustion process gives a

Page 253: Advanced Thermodynamics for Engineers

Examples of significance of Kp 237

product energy of

UP(Tp) = (1 - a ) x 276 960 + 43 993 = 279 575 kJ

This energy is contained in the products as shown in the following table:

Constituent CO, CO 0, N2

’2949 135 182 75 820 80 205 75 040 n 0.8506 0.2494 0.0747 1.88

Up(T,) = 280 961 kJ

The equations are balanced to within O S % , and this is close enough for this example.

Alternative method

In this approach the degree of dissociation, a , will not be introduced explicitly. The chemical equation which was written in terms of a in eqn (12.77) can be written as

1 1.1CO + A (Q + 3.76N2) -aCO, + bCO + c 0 2 + d N 2

n p = a + b + c + d 2 n R = 3.48

(12.86)

Considering the atomic balances,

Carbon: l . l = a + b giving b = 1.1 - a (12.87a) Oxygen: 2.1 = 2 a + b + 2c giving c = 0.5(1 - a ) (12.87b) Nitrogen: 1.88 = d (12.87~)

Total amount of substance in the products,

np= a + b + c + d (12.88)

The ratios of the amounts of substance in the equilibrium products are defined by the equilibrium constant

Hence

b2c - pp K ; r = a 2 Po nP

and, from the atomic balances, this can be written in terms of a as

(1.1 - a ) 2 1 - (1 - a ) - p P K i r = a 2 2 Po nP

(12.89)

(12.90)

(12.91)

The previous calculations showed that the temperature of the products which satisfies the governing equations is Tp = 2950 K, which gives a value of KZr = 13.1488. These values

Page 254: Advanced Thermodynamics for Engineers

238 Chernicai equilibrium and dissociation

will be used to demonstrate this example. Then

13.1488 { 1.21 -- 2.2a + a* - (1.21a + 2.2a2 + a 3 ) } pp x = a (12.92)

Po nP 2

It is possible to evaluate the ratio pp/pon, from the perfect gas relationship, giving

p p pPTP 85x2950

ponP nRTR 3 . 4 8 ~ 6 0 0 = 12.009 -- - -- -

This enables a cubic equation in a to be obtained:

95.532 -- 269.23~ + 251.65~' - 7 8 . 9 5 3 ~ ~ = 0 (12.93)

The solution to this equation is a = 0.8506, and hence the chemical equation becomes

1

2 1.lCO + - (02 + 3.76N,)-+0.8506CO2 + 0.2494CO + 0.074702 + 1.88N2

(12.94)

which is the same as that obtained previously. The advantage of this approach is that it is possible to derive a set of simultaneous equations which define the equilibrium state, and these can be easily solved by a computer program. The disadvantage is that it is not possible to use the intuition that most engineers can adopt to simplify the solution technique. It must be recognised that the full range of iteration was not used in this demonstration, and the solutions obtained from the original method were simply used for the first 'iteration'.

12.8 The Wan't Hoff relationship between equilibrium constant and heat of reaction

It has been shown that (eqn (12.59))

Thus

Consider the definition of p0,

p0 = ho + h(T) - T {so + s(T)] dh

= h, + h(T) - T s , - T I - T

Consider the terms

dh

T h(T)-T/ - = T

(12.95)

(12.96)

(12.97)

(12.98)

Page 255: Advanced Thermodynamics for Engineers

The effect of pressure and temperature on degree of dissociation 239

Let v = h(T),

1

dv = dg

dT d u = -- u = - T ' T2

Integrating by parts gives

(12.99)

( 1 2.1 00)

Although C has been used as a shorthand form it does include both +ve and -ve signs; these must be taken into account when evaluating the significance of the term. Thus

d 1

dT %T2 - ( h K p r ) = - - (y, h, + Vb hb - yc h, - vdhd)

d QP dT %T2 - ( InKpr) = -

(12.101)

(12.102)

( 12.103)

Equation (12.103) is known as the Van't Hoff equation. It is useful for evaluating the heat of reaction for any particular reaction because

The value of d(ln Kp)/d(l/T) may be obtained by plotting a graph of In Kp against 1/T. The values of Kp have been calculated using eqn (12.95), and are listed in Table 12.3, at

the end of the chapter, for four reactions. (The values have also been depicted as a graph in Fig 12.6, and it can be seen that over a small range of temperature, In Kp = A - b/T, where T is the temperature in K: this is to be expected from eqn (12.95)).

12.9 The effect of pressure and temperature on degree of dissociation

12.9.1 The effect of pressure

The effect of pressure on the degree of dissociation is defined by eqn (12.62), namely,

Page 256: Advanced Thermodynamics for Engineers

240 Chemical equilibrium and dissociation

It can be seen that the ratio of the amounts of substance is given by

( 1 2.1 04)

This can be interpreted in the following way. If v, + vd - v, - vb = 0 then the mole frac- tions of the products will not be a function of pressure. However, if v, + v d - v, - v b > 0 then the species on the left-hand side of the chemical equation (i.e. the reactants) increase, whereas if Y, + vd - v, - v,, < 0 then the species on the right-hand side of the equation (i.e. the products) increase. The basic rule is that the effect of increasing the pressure is to shift the equilibrium to reduce the total amount of substance.

Considering the two reactions introduced previously. The equation for the carbon monoxide reaction (12.27) was

1 a

2 2 CO + - Q - ( 1 - a)CO, + aCO + - 0,

and the equilibrium equation (12.72) was

_ _ _ ~ 1-a 1 + a / 2 / y l & - - 1 ; a / y ; ~-

1 + a / 2 a KPr =

This means that v, + vd - v, - v b < 0, and this would result in the constituents on the ‘products’ side of the equation increasing. This is in agreement with the previous statement because the total amounts of reactants in eqn (12.27) is 1.5, whilst the total amounts of products is 1 + a/2, where a is less than 1.0.

A similar calculation for the combustion and dissociation of a stoichiometric methane ( C h ) and air mixture, performed using a computer program entitled EQUIL2 gave the results in Table 12.1. The chemical equation for this reaction is

CH, + 2(0, + 3.76N2) 3 (1 - a,)CO, + a,CO + 2(1- a,)H,O + 2a,H2 + (a , /2 + a,)O, + 7.52N2

(1 2.105)

where a , is the degree of dissociation of the CO, reaction and a2 is the degree of dissociation of the H,O reaction. It can be seen that dissociation tends to increase the

Table 12.1 Amount of products for constant pressure combustion of methane in air. Initial temperature = 1000 K; equivalence ratio = 1.00

_______

No Pressure (bar) dissociation 1 10 100

Amount of C 0 2 1 0.6823 0.7829 0.8665 Amount of CO 0 0.3170 0.2170 0.1335 Amount of H,O 2 1.8654 1.920 1.9558 Amount of H, 0 0.1343 0.0801 0.0444 Amount of 0, 0 0.2258 0.1485 0.0889 Amount of N, 7.52 7.52 7.52 7.52 Total amount of substance 10.52 10.7448 10.6685 10.6091

Page 257: Advanced Thermodynamics for Engineers

The effect of pressure and temperature on degree of dissociation 241

amount of substance of products, and hence the effect of an increase in pressure should be to reduce the degree of dissociation. This effect can be seen quite clearly in Table 12.1, where the amount of substance of products is compared under four sets of conditions: no dissociation, dissociation at p = 1 bar, dissociation at p = 10 bar, and dissociation at p = 100 bar. The minimum total amount of substance occurs when there is no dissociation, while the maximum amount of substance occurs at the lowest pressure. Figure 12.4 shows how the degrees of dissociation for the carbon dioxide and water reactions vary with pressure.

12.9.2 THE EFFECT OF TEMPERATURE

The effect of temperature can be considered in a similar way to the effect of pressure. Basically it should be remembered that the changes in composition that take place during dissociation do so to achieve the minimum value of Gibbs energy for the mixture. The Gibbs energy of each constituent is made up of three components, the Gibbs energy at absolute zero (gn), the Gibbs energy as a function of temperature ( g ( T ) ) , and that related to partial pressure (see Fig 12.2). The equilibrium point is achieved when the sum of these values for all the constituents is a minimum. This means that, as the temperature rises, the constituents with the most positive heats of formation are favoured. These constituents include 0, (go = 0), H2 (go = 0), and CO (go = - 113 MJ/kmol). Both water and carbon dioxide have larger (negative) values of g,. This effect can be seen from the results in Table 12.2, whch have been calculated for the combustion of methane in air. Another way of considering this effect is simply to study eqn (12.55), and to realise that for gases with negative heats of formation an increase in temperature leads to a decrease in the value of K, . This means that the numerator of eqn (12.55) must get smaller relative to the denominator, which pushes the reaction backwards towards the reactants. This is borne out in Table 12.2: the degree of dissociation for this reaction increases with temperature, and this is shown in Fig 12.5.

Page 258: Advanced Thermodynamics for Engineers

242 Chemical equilibrium and dissociation

Table 12.2 Amount of products for constant pressure combustion of methane in air. Pressure = 1 bar; equivalence ratio = 1

Temperature (K)

Amount of CO, Amount of CO Amount of H,O Amount of H, Amount of 0, Amount of N, Total amount of substance

No dissociation T=1000K T=1500K T=2000K

1 0 2 0 0 7.52

10.52

0.6823 0.48 14 0.3031 0.3170 0.5186 0.6969 1.8654 1.7320 1.5258 0.1343 0.2679 0.474 1 0.2258 0.3933 0.5855 7.52 7.52 7.52

10.7448 10.9132 11.1055

0 7 -

0 6 -

E 0 5

m

- 0 .- * .- 4 0 4 - Q v) .- c

0 3 - z 2

0 2 -

0 1 -

0 + I I I 1 I I I I I 1

1000 I100 1200 1300 1400 1500 I600 1700 1800 1900 2000

Temperature / (K)

Fig. 12.5 Effect of temperature on degree of dissociation

-+-- Carbon dioxide - Water

Finally, it should be noted that in all the cases shown the degree of dissociation in the hydrogen reaction is much less than that for the carbon reaction. This supports the assumption made in previous work that the hydrogen will be favoured in the oxidation process.

12.10 Dissociation calculations for the evaluation of nitric oxide If it is necessary to evaluate the formation of nitric oxide (NO) in a combustion chamber then the equations have to be extended to include many more species. While it is possible to add the calculation of NO to a simple dissociation problem, as is done in Example 5 below, this does not result in an accurate estimate of the quantity of NO formed. The reason for this is that NO is formed by a chain of reactions that are more complex than simply

N, + 0, = 2 N 0

This chain of reactions has to include the formation of atomic oxygen and nitrogen, and

Page 259: Advanced Thermodynamics for Engineers

The effect of pressure ana' temperature on degree of dissociation 243

also the OH radical. To obtain an accurate prediction of the NO concentration a total of 12 species has be considered:

[11 H20, E21 H2, [31 OH, [41 H, E51 N2, [61 NO, 171 N, [81 CO2, E91 COS [lo1 0 2 , E11 1 0, E121 Ar.

The numbers in [ ] will be used to identify the species in later equations.

equations: The formation and breakdown of these species are defined by the following set of

where b = xi/.,.

106

1 os

i o 4

1 o3

102

101

100

10-1

10-2

1 o - ~

0 500 1000 1500 2000 2500 3000 3500

( 1 2.1 06)

Temperature / (K)

Fig. 12.6 Variation of equilibrium constant, Kh, at a standard pressure, p o = 1 bar, with temperature for the reactions CO + 0, e CO,, H2 + f 0, w HzO, CO, + H2 w CO + H,O, f N, + 10, NO.

Page 260: Advanced Thermodynamics for Engineers

Mol

fr

acti

on

Mol

fr

acti

on

- 0, N

Mol

fr

acti

on

Page 261: Advanced Thermodynamics for Engineers

Dissociation problems with two, or more, degrees of dissociation 245

A numerical method for solving these equations is given in Horlock and Winterbone (1986), based on the original paper by Lavoie et al. (1970).

Figure 12.7 shows the results of performing such calculations using a simple computer program. The coefficients in the program were evaluated using the data presented in Table 9.3, but with the addition of data for OH and N. The three diagrams are based on combustion of octane in air at a constant pressure of 30 bar, and show the effect of varying equivalence ratio, I $ , at different temperatures. The conditions are the same as quoted in Heywood (1988). Figure 12.7(a) shows the results for the lowest temperature of 1750 K, and it can be seen that the graph contains no atomic nitrogen (N) or atomic oxygen (0) because these are at very low concentrations When the mixture is weak some OH is produced. It is apparent from this diagram that there is not much dissociation of the carbon dioxide or water because the concentration of oxygen drops to very low values above I$ = 1. Some nitric oxide is formed in the weak region (xNo = 3 x at the weakest mixture), but this rapidly reduces to a very low value at stoichiometric simply because there is no oxygen available to combine with the nitrogen. Figure 12.7(b) shows similar graphs for a temperature of 2250 K, and it can be seen that the NO level has increased by almost a factor of 10. There is also some NO formed by combustion with rich mixtures; this is because, by this temperature, the carbon dioxide and water are dissociating, as is indicated by the increase in the CO and the OH radical in the weak mixture region. By the time 2750 K is reached, shown in Fig 12.7(c), there is a further tripling of the NO production, and CO is prevalent throughout the weak mixture zone. There are also significant amounts of OH and oxygen over the whole range of equivalence ratio.

The diagrams shown in Fig 12.7 are relatively contrived because they do not depict real combustion situations. However, they do allow the parameters which control the production of the products of combustion to be decoupled, to show the effects of changing the parameters independently of each other. The equilibrium concentrations depicted in Fig 12.7 are the values which drive the formation of the exhaust constituents through the chemical kinetics equations, which will be discussed in Chapter 14.

12.1 1 Dissociation problems with two, or more, degrees of dissociation

The previous example, which considered the dissociation of carbon monoxide, shows the fundamental techniques involved in calculating dissociation but is an unrealistic example because rarely are single component fuels burned. Even if a single component fuel such as hydrogen was burned in an engine, it would be necessary to consider other ‘dissociation’ reactions because it is likely that nitric oxide (NO) will be formed from the combination of the oxygen and nitrogen in the combustion chamber. These more complex examples will be considered here.

12.11.1 EXAMPLE 4: COMBUSTION OF A TYPICAL HYDROCARBON FUEL

A weak mixture of octane (C,H,,) and air, with an equivalence ratio of 0.9, is ignited at 10 bar and 500 K and bums at constant volume. Assuming the combustion is adiabatic, calculate the conditions at the end of combustion allowing for dissociation of the carbon dioxide and water, but neglecting any formation of NO.

Page 262: Advanced Thermodynamics for Engineers

246 Chemical equilibrium and dissociation

Solution

Stoichiometric equation neglecting dissociation

C8Hls + 12.5(02 + 3.76N2) SCO, + 9H,O + 47 N, nR = 60.5 np=64

Weak mixture neglecting dissociation

(12.107)

C8H18 + 13.889(02 + 3.76N2) 3 K O , + 9H@ + 1.38900, + 52.222N2 nR = 67.1 1 1 np = 70.61 1

( 12.108) Dissociation equations (12.25 and 12.34):

a l C O , ~ a l C O + - 0, a1 2

a2

2 a , H , 0 ~ a , H , + - 0,

Weak mixture including dissociation is

C8H,, + 13.889(0, + 3.76N2) nR = 67.1 I 1

8(1 - a1)C02 + 9(1 - a,)H,O + 8 a l C 0

2 8

0, + 52.222N2 (12.109)

np= 70.611 +4a, +4.5a2

The temperature and pressure calculated for combustion without dissociation are 2874 K and 60.481 bar respectively. Dissociation will lower the final temperature: assume that Tp = 2800 K.

The equilibrium constants from Table 12.3 at 2800 K are

Prco,

Prco Pro, = 6.58152 KP,, = 112

Prco PrH*O KPd = = 6.8295

Prco, Prnz

From eqn (12.109), the partial pressures of the constituents are

8(1 -a11 - P Prco, =

nP Po 9(1 -a,) - P Prn,o =

Sa, P np PO np PO 9a2 P

P r H z = - - Prco = - -

(1.389 + 4a1 + 4 . 5 ~ ~ ~ ) - p nP Po Pro, = nP Po

(12.110)

Page 263: Advanced Thermodynamics for Engineers

Hence

Po [8a1]* (1.389+4a1+4.5~2) p

- [8(1 - aJI2 nP Ki,, = 43.3164 =

Dissociation problems with two, or more, degrees of dissociation 247

(12.111)

and

(1 2.1 12)

The Kpr equation (eqn (12.1 11)) contains a pressure term, but .,.is can be replaced by the temperature of the products by applying the perfect gas law to the mixture. Then

PPVP = nP8TP

PRVR = n 8 T R Thus

nP nR3& nR&

P P PRSTP PRTP -=--- -

Substituting gives

(1 - all2 nR T R K;c, = 43.3164 = Po

a: (1.389 + 4Ql + 4 . 5 Q ~ ) p ~ T p

Inserting values for these parameters gives

- (1 - ah2 - 43.3164 x 10 x 2800

67.111 x 500 a:(1.389 + 4 a l + 4.5~12)

which can be expanded to give

(12.113)

(1 2.1 14)

(12.1 15)

(12.116)

144.58a:+ a:(49.2052 + 162.652~1,) + 2 a , - 1 = 0 (12.117)

This equation contains both a1 and a2. If it is assumed that a1 >> a2 then eqn (12.117) is a cubic equation in a l . Based on this assumption the value of a1 is approximately 0.11. Substituting this value in eqn (12.112) gives a2 = 0.017776, which vindicates the original assumption. Recalculation around this loop gives

a , = 0.108 a2 = 0.01742

These values are close enough solutions for the dissociation coefficients, and for the chemical equation (eqn (12.109)) which becomes

C,H,, + 13.889(0, + 3.76N2) nR=67.111

+ 7.136CQ + 8.84322 H,O + 0.846CO + 0.15678 H, + 1.89940, + 52.222N2 . - np=71.121

(12.11 8)

Page 264: Advanced Thermodynamics for Engineers

248 Chemical equilibrium and dissociation

This equation satisfies the equilibrium constraints and the chemistry of the problem, but it still has to be checked to see if it meets the First Law. The First Law is defined by the equation

( 1 2.1 1 9)

Note two things about this equation. First, that it is more convenient to evaluate the difference between the internal energies of the products, because now the composition of the products is also a function of temperature and hence Up(Ts) is not a constant. Second, that with dissociation, (Q,), is not the full value of the internal energy of reaction of octane because not all the octane has been oxidised fully to CO, and water. The value of ( QV)? in this case is given by

(Q,),= ( Q , ) , - O . l 0 8 ~ 8 ~ (-282 9 9 3 ) - 0 . 0 1 7 4 2 ~ 9 ~ (-241 827) = -5 116 320 + 244 505.9 + 37 913.6 = -4 833 900.5 kJ/kmol octane (1 2.120)

Alternative method for calculating the energy released by combustion

The energy released from partial combustion can be calculated using Hess' law. In this case

Products Reactants

= 7.136(AUo)oz + 8.843 22(AUo)H20 + 0.846(AU,),-- - (AUO)CxHIx = 7 . 1 3 6 ~ (-393 405)+8.84322~ (-239 082)

+ 0.846 x (- 113 882) - (-74 897)* = -4 943 040 W/kmol C,H,, burned (12.121)

This value can be related to the energy released at the standard temperature of 25°C by the following equation (*note, the fuel properties have been based on methane):

- ( Q v ) s = + [Up(Ts) - UP(TO)I - [UR(TS) - UR(TO)I

= -4 943 040 + 7.136 x 7041.6 + 8.84322 x 7223.3 + 0.846 x 601 1.8 + 0.15678 x 6043.1 + 1.8994 x 6032.0 + 52.222 x 6025.8 - (5779 + 13.889 x [6032.0 + 3.76 x 6025.81)

= -4 895 205 kJ/kmol C,H,, burned (12.122)

This is within 1.25% of that calculated from the internal energy of reaction, in eqn (12.120), and is probably because of the approximation made for the properties of octane.

Evaluating the energy terms for the reactants gives

Constituent C,H,, 02 N*

U5(x) 45 783 10 574.8 10 356.1 U29x 0 6032.0 6025.8 Difference 45 783 4542.8 4330.3 n 1 .o 13.889 52.222

Page 265: Advanced Thermodynamics for Engineers

Dissociation problems with two, or more, degrees of dissociation 249

Hence, the energy of the products, based on the evaluated degrees of dissociation, must be

U P V P ) - UP(T,) = - ( & A + EU,(TR) - U,(Ts)I = 4 833 716.2 + 335 014 = 516 8731 W

Evaluating the energy contained in the products at 2800 K gives

(12.123)

Constituent CO, H2O co H2 0 2 N2 _ _ _ _ _ _ ~ ~ ~ ~ ~

u28W 127 154.9 100 470.8 71 512.4 66 383.6 75 548.1 70 754.1 u29X 7041.6 7223.3 6011.8 6043.1 6032.0 6025.8 Difference 120 113.3 93 247.5 65 500.5 60 340.5 69 516.1 64 728.4 n 7.1360 8.84322 0.864 0.15678 1.8994 52.222

Up(Tp) - Up(T,) = 5 260 072.5 kJ This shows that the products at 2800 K contain more energy than was available from the

energy released by the fuel and hence the energy equation has not been satisfied. It is necessary to repeat the whole calculation with a lower temperature guessed for the products. Rather than do a number of iterations, the value obtained from a computer program will be used immediately, and it will be shown that this gives good agreement in the energy equation: Tp = 2772 K will be used.

At Tp = 2772 K the values of equilibrium constant can be obtained from the tables by linear interpolation:

22

50 22

50

Kpr, = 8.14828 + - x (6.58152 - 8.14828) = 7.4589 ( 1 2.1 24)

KPrZ = 6.72829 + - x (6.82925 - 6.72829) = 6.7727

Substituting these values gives

55.6352 x 10 x 2772 - - (1 -all2 67.111 x 500 a:(1.389 + 4 a l + 4 . 5 ~ 1 ~ )

and hence

183.84a:+ a:(62.838 + 206.82a,) + 2 a , - 1 = 0

(1 2.125)

( 12.126) Assuming a2 = 0 for the initial iteration gives a1 = 0.099. Recalculation around the loop results in the values

a , = 0.097 a2 = 0.0156

which give a chemical equation of

(1 2.127)

j 7.224CQ + 8.8596 H,O + 0.776 CO + 0.1404 H, + 1.8472 0, + 52.222 N, np=71.061

(12.128)

Page 266: Advanced Thermodynamics for Engineers

250 Chemical equilibrium and dissociation

The energy released by the fuel becomes

(Q,), = (Q,), - 0.097 x 8 x (-282 993) - 0.0156 x 9 x (-241 827) = -5 116 320 + 219 602.7 + 33 952.5 = -4 862 764.8 kJ/kmol octane (1 2.129)

Hence, the energy of the products, based on the evaluated degrees of dissociation, must be

UP(TP) - UP(',)= - ( Q v > s + [UR(TR) - uR(Ts)I

= 4 862 764.8 + 335 014 = 5 197 778.8 kJ

Evaluating the energy contained in the products at 2772 K gives

(12.130)

Constituent CO, H20 co H2 0 2 N,

112772 125 654.9 99 196.2 70 702.8 65 594.3 74 678.0 69 948.5 u298 7041.6 7223.3 6011.8 6043.1 6032.0 6025.8 Difference 118 613.3 91 972.9 64 690.8 59 551.2 68 646.0 63 922.7 n 7.2240 8.8596 0.776 0.1404 1.8472 52.222

Up(Tp) - Up(T,) = 5 195 240 kl

These values are within 0.05%, and hence satisfy the energy equation.

Alternative method for calculating the chemical equation

The approach used above to calculate the coefficients in the chemical equation was based on the degrees of dissociation of the two reactions occurring in this example. It was shown previously that this approach is often the best for manual solution, but that a more general approach, in which a system of simultaneous equations is developed, is better for computer solution. This second method will be outlined below, for one of the steps in the previous example.

First, eqn (1 2.109) can be replaced by C,H,, + 13.889(02 + 3.76N2) - aCO, + bH,O + c C 0 + dH, + eo, +fN,

nR = 67.1 11 np = a + b + c + d + e + f

(12.13 1) If the first iteration of the previous approach is used then Tp = 2800 K and, as given in eqn ( 12.1 lo),

P I C O ,

P I C 0 P r o ,

= 6.58152 KPr, = 112

P r c o P*H,O

P I C O , P I H ,

KPd = = 6.8295

Now, by definition,

P r c o , a nY2 112 =- KP,, = 112 cell, py' Po PIC0 P r o ,

(1 2.132)

Page 267: Advanced Thermodynamics for Engineers

Dissociation problems with two, or more, degrees of dissociation 251

which gives

For the water gas reaction

Prco PIH,O cb

ad Prcoz Prnz = 6.8295 = - KP,z =

(12.133)

(12.134)

and the total amount of substance in the products is

np= a + b + c + d + e + f (12.135)

Considering the atom balances gives

Carbon: 8 = a + c * c = 8 - a (1 2.136a)

Hydrogen: 18 = 2b + 2d * b = 9 - d (12.136b)

Oxygen: (12.136~)

Nitrogen: f = 52.223 (12.136d)

27.778 = 2a + b + c + 2e e = 13.889 - a - (b + c)/2

From the perfect gas law

np ~ R T R PP PRTP - =-

giving eqn (12.133) as 2

2 a n R T R Kp,l = 43.3164 = -y- - Po

c e PRTP

(1 2.137)

(12.138)

Equation (12.138) is a member of a non-linear set of equations in the coefficients of the chemical equation. One way to solve this is to make an assumption about the relative magnitude of d and the other parameters. If it is assumed that d is small compared with the other parameters in eqn (12.131) then eqn (12.138) becomes a cubic equation in a. This is equivalent to assuming that a2 is small compared with a , , as was assumed in the previous calculation. If eqns (12.136a, b and c) are substituted into eqn (12.138), with the assumption that d = 0, then eqn (12.138) becomes

a2 67.111 x 500 2

x l a n R T R (8 - ~)’(5.389 - ~ / 2 ) = 7 - Po =

K ~ r l PR T P 43.3164 10 x 2800

= 0.027667 a’ (12.1 39)

giving

344.896 - 118.224~ + 13.3613~’ - a3/2 = 0 (12.140)

The solution to eqn (12.140) relevant to this problem is a=7.125, giving c = 8 - a=0.875 and b = 9 . It is now possible to check if this result satisfies the

Page 268: Advanced Thermodynamics for Engineers

252 Chemical equilibrium and dissociation

equilibrium condition for the water gas reaction, namely eqn (12.134), which gives

ch 0.875 x 9 = 0.16184 d = ____ = __

aKp12 7.125 x 6.89295 (12.141)

The calculation should now be performed again based on the new value of d given in eqn (12.141). This produces a new cubic equation in a , similar to that in eqn (12.140) but with the following coefficients:

350.07 - 119.519~ + 1 3 . 4 4 2 ~ ~ 2 -a3/2 = 0

The new value of a is 7.134, which is close enough to the actual solution because the value of d is compatible with this is

cb 0.866 x 8.8382

aKPrZ 7.134 x 6.89295 d = ~ - - = 0.1556 (1 2.142)

and it can be seen that the change in d will have a small effect on subsequent values of the coefficients. Hence, the chemical equation is

C,H,, + 13.889(0, + 3.76N2)

a 7.134CQ + 8.8382H20 + 0.866CO + 0.1618H2 + 1.9029002 + 52.222N2

n R = 67.1 11

np = 7 1,125

(12.143)

Equation (12.143), which has been calculated from the general equation [ eqn (12.131)), is almost the same as that calculated using the degrees of dissociation in eqn (12.109). This is to be expected because both approaches are exactly equivalent. This solution requires further iterations to satisfy the energy equation, as done in the previous approach.

12.11.2

A rich mixture of octane (C,H,,) and air, with an equivalence ratio of 1.1, is ignited at 10 bar and 500 K and bums at constant volume. Assuming the combustion is adiabatic, calculate the conditions at the end of combustion allowing for dissociation of the carbon dioxide and water, but neglecting any formation of NO.

EXAMPLE 5: A RICH MIXTURE

Solution

Stoichiometric equation neglecting dissociation:

C,H,, -+ 12.5 (Q + 3.76 N2) 3 8CQ + 9H,O + 47 N, tIR = 60.5 n p = 6 4

Rich mixture neglecting dissociation:

( 12.144)

C,H,, t 11.36(Q + 3.76NJ 3 5.7272CQ + 9H,O + 2.2728CO + 42.71 N2 n R =55 .09 np = 59.73

(12.145)

Page 269: Advanced Thermodynamics for Engineers

Dissociation problems with two, or more, degrees of dissociation 253

Dissociation equations (12.25 and 12.34):

a1

2 U , C O ~ ~ U , C O + - 0,

a2

2 U ~ H ~ O - U Z H , + - 0,

Rich mixture including dissociation is

C8H18 = 11.36(0, + 3.76N2)

3 5.7272(1 - u 1)COZ + 9(1 - u,)H,O + (2.2728 + 5.7272 u l)CO

0, + 42.71 N, (12.146) 2

5.7272

np=59.73+2.8636a, +4.5a2

It can be seen that this is a significantly more complex equation than for the weak mixture. Application of a computer program gives the final temperature and pressure, without dissociation, as 2958.6 K and 64.15 bar respectively. The equilibrium constants are

Prco,

Prco Proz

Prco PrH,O

112 Kpr, = 6.581 52 =

KPr2 = 6.829 25 = Prco2 Prn2

Substituting for the mole fractions gives

5.7272(1 - a l ) (2.2728 + 5 . 7 2 7 2 ~ ~ ) Pco, = P ; p c o = P

np nP ( 2 . 8 6 3 6 ~ ~ + 4.5~2,)

n P Po, = P

Hence

Po (2.2728 + 5.7272~1)’(2.8636~1 +4.5U2) p

- 5.7272,(1 - u J 2 n p K i r , = 43.3164 =

Similarly

and

(1 2.147)

( 12.148)

(1 2.149)

( 12.150)

(2.2728 + 5 . 7 2 7 2 ~ ~ ) x 9 x (1 - ~ 2 )

9U2 x 5.7272(1 - ~ l )

1 - ~2

a2

(2.2728 + 5.7272~1)

5.7272( 1 - (X 1)

X -- - KPrz =

(12.15 1)

Page 270: Advanced Thermodynamics for Engineers

254 Chemical equilibrium and dissociation

The K,, and Kh2 equations (eqns (12.149) and (12.151)) are again non-linear equations in a , and az , and they can be solved in the same manner as before.

The solution to this problem, allowing for dissociation is a final temperature of 2891 K and a final pressure of 62.99 bar. These result in the following degrees of dissoci- ation: a1 = 0.0096, az = 0.05540. It is left to the reader to prove these values are correct.

12.11.3

When a mixture of oxygen and nitrogen is heated above about 2000 K these two elements will join together to form nitric oxide (NO). The elements combine in this way because the energy of formation of nitric oxide is positive, i.e. the reaction which forms it is endothermic, and hence a combination of nitrogen and oxygen tends to reduce the Gibbs energy of the mixture. Note that other compounds tend to dissociate at high temperature because their energies of formation are negative, i.e. the reactions which form them are exothermic.

EXAMPLE 6: THE FORMATION OF NITRIC OXIDE

Examp le

Methane is burned with a stoichiometric quantity of air and achieves a pressure of 99.82 bar and a temperature of 2957 K after combustion, if dissociation of the carbon dioxide and water vapour are taken into account. The mole fractions of the constituents are 7.495% CO,; 1.890% CO; 18.183% H,O; 0.6425% H,; 1.267% 0,; and 70.58% N,. Calculate the amount of nitric oxide (NO) formed at this temperature, neglecting the effect of the NO formation on the dissociation of carbon monoxide and water. Estimate the effect of the NO formation on the temperature of the products. The energy of the reactants, U,(T,) = 179 377 kJ.

Solution:

Chemical equation without dissociation:

CH, + 2(0 , + 3.76N,) 3 CO, + 2H,O + 7.52N2 (12.152)

Chemical equation with dissociation:

C H, + 2 ( 0 2 + 3.76N2)

3 np(0.07495 C02 + 0.0189 CO + 0.18183 H,O + 0.006425 H, + 0.01267 O2 + 0.7058 N2) (12.153)

( I o np = 10.52+ 1 + 3

2 2

I t is possible to evaluate the total amount of substance in the products as follows:

1 - a I a1

nP np 0.07495 = - , and 0.0189 = - 3 np = 10.655

Page 271: Advanced Thermodynamics for Engineers

Dissociation problems with two, or more, degrees of dissociation 255

and hence

CH, + 2(0, + 3.76N2) - 0.7986 CO, + 0.2014CO + 1.9315H20 + 0.06845H2 + 0.134950, + 7.52N2 (12.154)

np = 10.655

Assume that combination of the nitrogen and oxygen occurs to form nitric oxide (NO):

a3 a3

2 2 a3NO= - 0, + - N,

The temperature of the products is TP = 2956 K, giving

PrNO 112 112

Pro? PIN^ ICpr3 = 0.1 1279 =

(12.155)

(1 2.156)

Adding in the NO equation to the chemical equation gives

CH, + 2(0, + 3.76N2)

3 0.7986 CO, + 0.2014 CO + 1.9315 H,O + 0.06845 H,

+ a3N0 - + 0.13495 - - 0, + 7.52 - - N, (12.157) ( ";i ( ;) n p = 10.655

The partial pressures of the nitrogen constituents are

(7.52 - a3/2) (0.13495 - a3/2) p (12.158) a3

nP nP nP P N O = - P ; PN, = P ; Po,=

and hence

= 0.1 1279 a3

(7.52 - a3/2)112(0.13495 - a3/2)112

Squaring gives

1.00318a;+ 0.048690a3 - 0.01291 = 0

which gives

a3 = 0.09174

Hence, the chemical equation is

CH, + 2(0 , + 3.76N2)

(1 2.159)

( 12.160)

* 0.7986C0, + 0.2014CO + 1.9315H20 + 0.06845H2 + 0.09174NO + 0.08908 0, + 7.474N2 (12.161)

np = 10.655

Page 272: Advanced Thermodynamics for Engineers

256 Chemical equilibrium and dissociation

It can be seen that the effect of the NO formation is to reduce the amount of oxygen in the products. The effect of this is to upset the equilibrium of the carbon dioxide and water reactions, and the values of a , and a2 will be changed. This will be considered later. First, the effect of the formation on the energy equation will be considered. This effect has already been shown in Fig 12.7, but a more complex set of reactions was used. It is not possible to demonstrate the more complex calculation in a hand calculation.

The energy released by the original reaction, i.e. neglecting the NO reaction, can be evaluated by considering Hess' equation:

AUo = Uf - 1 Ui products Reactants

=0.7986 x (-393 405) + 0.2014 x (-113 882) + 1.9315 x (-239 082) -(-66 930)

= -73 1 966 kJ/kmol CH,

( 1 2.1 62)

The effect of the formation of NO is to reduce AUo in the following manner:

Products Reactants

~ 0 . 7 9 8 6 x (-393 405) + 0.2014 x (-113 882) + 1.9315 x (-239 082) +0.09174 x 89 915, - (-66 930)

Formation of NO

= -723 717kJ/km01 CH, (12.163)

The energy equation based on 0 K is

- A u O + uR(0)l - [uP(TP) - = O (12.164)

giving

UdTP) = -Auo + u R ( T R ) (12.165)

Applying the energy equation to the chemical equation neglecting dissociation of NO (i.e. eqn (1 2.154)) gives

(1 2.166) Up(Tp)=-(-731 927)+ 179 377=911 304k.J

Substituting the value of T, given in the question namely 2957 K, gives

Constituent CO, H,O co H2 0, N,

u2957 135 615 107 643 76050 70837 80457 75 270 n 0.7986 1.9315 0.2014 0.06845 0.13495 7.52

Up(Tp) = 913 268 kJ

This satisfies the energy equation within 0.2%.

Page 273: Advanced Thermodynamics for Engineers

Dissociation problems with two, or more, degrees of dissociation 257

Now considering the case with NO formation (Le. in eqn (12.157)). First, the energy of the products is less than in the previous case because the formation of NO has reduced the energy released by combustion. In this case

Up(Tp) = -( -723 678) + 179 377 = 903 055 kJ

If it is assumed that the temperature is reduced by the ratio 903 055/911 304, then a first approximation of the products temperature is 2930 K. In addition, it is necessary to include NO in the products, and this is introduced in the table below:

Constituent CO, H,O CO H2 0, NO NZ

'2930 134 153 106407 75 271 70068 79609 77 304 74494 n 0.7986 1.9315 0.2014 0.06845 0.08908 0.09174 7.474

UP(Tp) = 903 567 kJ

This value of the products' energy obviously satisfies the energy equation, and shows that the effect of the formation of NO has been to reduce the products temperature, in this case by about 30 K.

Table 12.3 Equilibrium constants

500 550 600 650 700 750 800 850 900 950

1000

1050 1100 1150 1200 1250 1300 1350 1400 1450 1500

1.06909E + 25 2.16076E + 22 1.22648E + 20 1.54176E + 18 3.62178E + 16 1.40361E + 15 8.17249E + 13 6.65502E + 12 7.16841E + 11 9.77343E + 10 1.62821E + 10

3.22109E + 09 7.39213E + 08 1.93024E + 08 5.64323E + 07 1.82234E + 07 6.42591E + 06 2.45016E + 06 1.00177E + 06 4.38021E + 05 2.00768E + 05

7.90482E + 22 3.78788E + 20 4.38630E + 18 1.00209E + 17 3.90700E + 15 2.33722E + 14 1.98055E + 13 2.23642E + 12 3.20880E + 11 5.63412E + 10 1.17472E+ 10

2.83842E + 09 7.79082E + 08 2.38939E + 08 8.07620E + 07 2.97385E + 07 1.18132E +07 5.02025E + 06 2.26609E + 06 1.07979E + 06 5.40233E + 05

0.00739 0.01753 0.03576 0.06500 0.10788 0.16651 0.24234 0.33605 0.44763 0.57647 0.72148

0.88 120 1.05393 1.23787 1.43113 1.63 189 1.83837 2.04895 2.26209 2.47646 2.69084

1.5710E - 09 1.1361E - 08 5.91 12E - 08 2.3872E - 07 7.8997E - 07 2.2291E - 06 5.5257E - 06 1.231 1E - 05 2.5096E - 05 4.7465E - 05 8.4233E - 05

1.4154E - 04 2.2687E - 04 3.4904E - 04 5.1805E-04 7.4501E - 04 1.0419E - 03 1.4212E - 03 1.8962E - 03 2.4800E - 03 3.1860E-03

(continued)

Page 274: Advanced Thermodynamics for Engineers

Table 12.3 (continued)

1550 1600 1650 1700 1750 1800 1850 1900 1950 2000

2050 2100 2150 2200 2250 2300 2350 2400 2450 2500

2550 2600 2650 2700 2750 2800 2850 2900 2950 3000

3050 3100 3150 3200 3250 3300 3350 3400 3450 3500

9.72637E + 04 4.93402E + 04 2.60986E + 04 1.4341 3E + 24 8.15991E + 03 4.79334E + 03 2.89945E + 03 1.80179E + 03 1.14787E + 03 7.48280E + 02

4.98289E + 02 3.38436E + 02 2.34122E + 02 1.64752E + 02 1.17799E + 02 8.54906E + 01 6.29132E + 01 4.69061E + 01 3.54023E + 01 2.70287E + 01

2.08601E + 01 1.62640E + 01 1.28029E + 01 1.01701E + 01 8.14828E + 00 6.58152E + 00 5.35700E + 00 4.39219E + 00 3.62613E + 00 3.01343E + 00

2.51997E + 00 2.1 1989E + 00 1.79349E + 00 1.52559E + 00 1.30444E + 00 1.12089E + 00 9.67731E - 01 8.39304E - 01 7.31095E - 01 6.39503E - 01

2.82471E + 05 1.53723E + 05 8.67587E + 04 5.06191E + 24 3.04452E + 04 1.88294E + 04 1.19480E + 04 7.76297E + 03 5.15523E + 03 3.49343E + 03

2.41215E + 03 1.69484E + 03 1.21032E + 03 8.77496E + 02 6.45249E + 02 4.80785E + 02 3.62704E + 02 2.76820E + 02 2.13590E + 02 1.66502E + 02

1.31054E + 02 1.04097E + 02 8.33978E + 01 6.73591E + 01 5.48240E + 01 4.49468E + 01 3.71035E + 01 3.08294E + 01 2.57753E + 01 2.89620E + 01

2.47329E + 01 2.12369E + 01 1.83304E + 01 1.59006E + 01 1.38587E + 01 1.21342E + 01 1.06707E + 01 9.42300E + 00 8.35460E + 00 7.43587E + 00

2.90418 3.11558 3.32426 3.52960 3.73107 3.92824 4.12080 4.30849 4.49113 4.66861

4.84087 5.00786 5.16961 5.32615 5.47753 5.62384 5.76515 5.90158 6.03322 6.160 17

6.28255 6.40045 6.51398 6.62323 6.72829 6.82925 6.92618 7.01 914 7.10821 7.19343

7.27483 7.35247 7.42636 7.49651 7.56295 7.62566 7.68464 7.73988 7.79136 7.8 3905

4.0274E - 03 5.0168E - 03 6.1666E - 03 7.4884E - 03 8.9931E - 03 1.0691E - 02 1.2590E - 02 1.4700E - 02 1.7028E - 02 1.9579E - 02

2.2359E - 02 2.5373E - 02 2.8624E - 02 3.21 15E - 02 3.5847E - 02 3.9822E - 02 4.4039E - 02 4.8498E - 02 5.3198E - 02 5.8137E - 02

6.3312E - 02 6.8721E - 02 7.4361E - 02 8.0227E - 02 8.6315E - 02 9.2620E - 02 9.9138E - 02 1.0586E - 01 1.1279E - 01 1.1992E - 01

1.2723E - 01 1.3473E - 01 1.4240E - 01 1.5025E - 01 1.5826E - 01 1.6643E - 01 1.7476E - 01 1.8322E - 01 1.9183E - 01 2.0057E - 01

Page 275: Advanced Thermodynamics for Engineers

Problems 259

12.12 Concluding remarks

It has been shown that dissociation is an equilibrium process which seeks to minimise the Gibbs energy of a mixture. In a combustion process, dissociation always reduces the temperatures and pressures achieved, and hence reduces the work output and efficiency of a device.

The thermodynamic theory of dissociation has been rigorously developed, and it will be seen later that this applies to fuel cells also (see Chapter 17).

Dissociation usually tends to break product molecules shown into intermediate reactant molecules (e.g. CO and H,). However, the same theory shows that at high temperatures, nitrogen and oxygen will combine to form oxides of nitrogen.

A wide range of examples has been presented, covering most situations which will be encountered in respect of combustion.

PROBLEMS

Assume that air consists of 79% N, and 21% 0, by volume.

1 A cylinder contains 1 kg carbon dioxide (CO,), and this is compressed adiabatically. Show the pressure, temperature and specific volume are related by the equations

3 dF = K,, /’- a 1.01325

and

%(2 + a)T 88

pv =

where a = degree of dissociation. Note that the equilibrium constant for the reaction CO + 0, w CO, is given by

(Pco,/Po)

(Pco / P o ) (Po,/Po)l~z KPr =

where po is the datum pressure of 1 atm. CO,, -391 032; H,O, -239 342; CO, -108 052 k.J/km~l.

2 (a) Calculate the equilibrium constant, K,, at 2000 K at the reference pressure of 1 atm for the reaction

CO + H,O CO, + H,

(b) Calculate the equilibrium constant, Kp, at 2500 K at the reference pressure of 1 atm for the reaction

1

2 co + - o,wco,

Also calculate the equilibrium constant at 2500 K at a pressure of 1 bar. [(a) 4.6686; (b) 27.028 atrn-”,; 27.206 bar-’/*]

Page 276: Advanced Thermodynamics for Engineers

260 Chemical equilibrium and dissociation

3 If it is assumed that the enthalpy of reaction, Q,, is a constant, show that the value of K , is given by

KP = e -QP’sT+k

where k is a constant.

Evaluate K, at the temperature of 2500 K. For a particular reaction Qp = -277 346 W/kmol, and K, = 748.7 at T = 2000 K.

[26.64]

4 A stoichiometric mixture of propane (CJ&) and air is burned in a constant volume bomb. The conditions just prior to combustion are 10 bar and 600 K. Evaluate the composition of the products allowing for dissociation if the peak temperature achieved is 2900 K, and show that the combustion process is adiabatic (i.e. the heat transfer is small).

The equilibrium constant for the CO, reaction at 2900 K is Kpr = pco2/pcopo2”2 = 4.50, and it can be assumed that the degree of dissociation for the water reaction is one-quarter of that for the carbon dioxide one. Indicate the processes, including a heat transfer process, on the U-T diagram. The internal energy of propane at 25°C is 12 380kJ/mol, and the internal energies of formation for the various compounds are CO,, -391 032; H20, -239 342; CO, - 108 052 U/kmol.

[0.0916; 0.0229; 0.1450; 0.0076; 0.01527; 0.71761

5 Methane (C&) is burned with 50% excess air. The equilibrium products at a pressure of 10 bar and a temperature of 1600 K contain CO,, CO, H,O, H,, 0, and N,. Calculate the partial pressures of these gases, assuming that the partial pressures of CO and H2 are small.

[0.6545; 1.6383 x 1.3089; 1.0524 x 0.6545; 7.38221

6 The exhaust gas of a furnace burning a hydrocarbon fuel in air is sampled and found to be 13.45% CO,; 1.04% CO; 2.58% 0,; 7.25% H,O; 75.68% N,; and a negligible amount of H2. If the temperature of the exhaust gas is 2500 K, calculate:

(a) (b) the equivalence ratio; (c)

(d)

the carbon/hydrogen ratio of the fuel;

the equilibrium constant for the dissociation of CO, (show this as a function of pressure); the pressure of the exhaust gas.

[1.0; 0.898; 8 0 . 5 2 / ~ ” ~ ; 8.8741

7 A weak mixture of propane (C31-&) and 50% excess air is ignited in a constant volume combustion chamber. The initial conditions were 1 bar and 300 K and the final composition was 7.87% CO,, 0.081% CO, 10.58% H,O, 0.02% H,, 6.67% 0, and 74.76% N, by volume. Evaluate the final temperature and pressure of the products. Prove that the temperature obeys the conservation of energy, if the calorific value of propane is 46 440 U/kg at 300 K.

The equilibrium constant for the carbon dioxide reaction is given by

(Pco)2Po, 28600 where log,, K , = 8.46 - - , and T = temperature in K.

T 2 K , =

( P C O J

Page 277: Advanced Thermodynamics for Engineers

Problems 261

The specific heats at constant volume (c,,,) of the constituents, in kJ/kmolK, may be taken as

co, 12.7 + 22 x 10-~77 H,O 22.7 + 8.3 x 10-3T

CO 19.7 + 2.5 x 10-3T N, 17.7 + 6.4 x 10-3T

0, 20.7 +4.4 x 10-3T H, 17.0 + 4.0 x 10-3T

C3H, neglect the internal energy in the reactants

The internal energies of formation of the constituents at 300 K are CO,, -393 x lo3;

[2250 K; 7.695 bar] H,O, -241 x 103; c o , - 112 x 103 k~/kmoi.

8 A mixture of propane and air with an equivalence ratio 0.9 (i.e. a weak mixture) is contained in a rigid vessel with a volume of 0.5 m3 at a pressure of 1 bar and 300 K. It is ignited, and after combustion the products are at a temperature of 2600 K and a pressure of 9.06 bar. Calculate:

(a) (b) (c)

the amount of substance in the reactants; the amount of substance in the products; the amounts of individual substances in the products per unit amount of sub- stance supplied in the fuel.

[0.02005; 0.02096; 0.0932; 0.1368; 0.01 138; 0.00261; 0.02792; 0.72811

9 A stoichiometric mixture of carbon monoxide and air reacts in a combustion chamber and forms exhaust products at 3000 K and 1 bar. If the products are in chemical equilibrium, but no reactions occur between the nitrogen and the oxygen, show that the molar fraction of carbon monoxide is approximately 0.169. The products of combustion now enter a heat exchanger where the temperature is reduced to 1000 K Calculate the heat transfer from the products per amount of substance of CO, in the gases leaving the heat exchanger if dissociation can be neglected at the lower temperature.

[ -399.3 MJ/kmol CO,]

10 A mixture containing hydrogen and oxygen in the ratio of 2 : 1 by volume is contained in a rigid vessel. This is ignited at 60°C and a pressure of 1 atm (1.013 bar), and after some time the temperature is 2227°C. Calculate the pressure and the molar composition of the mixture.

[5.13 bar, 96.5%; 2.36%; 1.18961

11 A vessel is filled with hydrogen and carbon dioxide in equal parts by volume, and the mixture is ignited. If the initial pressure and temperature are 2 bar and 60°C respectively, and the maximum pressure is 11.8 bar, estimate

(a) the maximum temperature; (b) (c)

the equilibrium constant, K p ; and the volumetric analysis of the products at the maximum temperature.

Page 278: Advanced Thermodynamics for Engineers

262 Chemical equilibrium and dissociation

Use

12

13

14

1354 T

logloKp = 1.3573 - -

Pn,oPco where K, = ~

PH2PC4 and T = temperature in kelvin

[1965; 4.6573; 0.3417; 0.3417; 0.1583; 0.15831

A stoichiometric mixture of hydrogen and air is compressed to 18.63 bar and 845°C. It bums adiabatically at constant volume. Show that the final equilibrium temperature is 3300 K and the degree of dissociation is 8.85%. Calculate the final pressure after combustion. Show the process, including the effect of dissociation, on a U-T diagram. The equilibrium constant is

(PH,O/PO)

(PHJPO) (Po,lPo)1/2 KPr = = 12.1389

at T = 3300 K

Use the following internal energies: Internal energy (k.J/kmol)

T (K) Oxygen, 0, Hydrogen, H2 Nitrogen, N, Water, H,O

0 0 0 0 -239 081.7 1100 25 700.3 23 065.3 24 341.8 1150 27 047.4 24 187.6 25 579.1 3300 91 379.8 80633.5 85 109.6 142 527.4

[47.58 bar]

A mixture containing equal volumes of carbon dioxide (CO,) and hydrogen (H,) is contained in a rigid vessel. It is ignited at 60°C and a pressure of 1 a m , and after some time the temperature is 2227°C. Calculate the pressure and the molar composition of the mixture.

[7.508 atm; 35.6%; 35.6%; 14.4%; 14.4%]

A gas turbine combustion chamber receives air at 6 bar and 500 K. It is fuelled using octane (C,H,,) at an equivalence ratio of 0.8 (Le. weak), which bums at constant pressure. The amount of carbon dioxide and water in the products is 9.993% and 1 1.37% by volume respectively. Calculate the maximum temperature achieved after combustion and evaluate the degrees of dissociation for each of the following reactions:

1 2 1 2

co + - 02-CO2

H, + - 02-H20

Page 279: Advanced Thermodynamics for Engineers

Problems 263

Assume the Octane is in liquid form in the reactants and neglect its enthalpy; use an

[0.0139; 0.00269; 2200 K] enthalpy of reaction of octane at 300 K of -44 820 kJ/kg.

15 A mixture of one part by volume of vaporised benzene to 50 parts by volume of air is ignited in a cylinder and adiabatic combustion ensues at constant volume. If the initial pressure and temperature of the mixture are 10 bar and 500 K respectively, calculate the maximum pressure and temperature after combustion. (See Chapter 10, Question 1.)

[5 1.67 bar, 2554 K]

16 A gas engine is operated on a stoichiometric mixture of methane ( C b ) and air. At the end of the compression stroke the pressure and temperature are 10 bar and 500 K respectively. If the combustion process is an adiabatic one at constant volume, calculate the maximum temperature and pressure reached. What has been the effect of dissociation? (See Chapter 10, Question 5).

[59.22 bar, 2960 K]

17 A gas injection system supplies a mixture of propane (C,&) and air to a spark ignition engine, in the ratio of volumes of 1 : 30. The mixture is trapped at 1 bar and 300 K, the volumetric compression ratio is 12:1, and the index of compression is 1c = 1.4. Calculate the equivalence ratio, the maximum pressure and temperature achieved during the cycle, and also the composition (by volume) of the dry exhaust gas. What has been the effect of dissociation? (See Chapter 10, Question 6.)

0.8463,0.1071,0.0465] [0.79334, 119.1 bar, 2883 K;

18 A 10% rich mixture of heptane (C,H,,) and air is trapped in the cylinder of an engine at a pressure of 1 bar and temperature of 400 K. This is compressed and ignited, and at a particular instant during the expansion stroke, when the volume is 20% the trapped volume, the pressure is 27.06 bar. Assuming that the mixture is in chemical equilibrium and contains only CO,, CO, H,O, H, and N,, find the temperature and molar fractions of the constituents. This solution presupposes that there is no 0, in the products; use your results to confirm this.

[2000 K; 0.1037; 0.0293; 0.1434; 0.00868; 0.71481

19 A turbocharged, intercooled compression ignition engine is operated on octane (C,H,,) and achieves constant pressure combustion. The volumetric compression ratio of the engine is 20 : 1, and the pressure and temperature at the start of compression are 1.5 bar and 350 K respectively. If the air-fuel ratio is 24 : 1 calculate maximum temperature and pressure achieved in the cycle taking into account dissociation of the carbon dioxide and water vapour. Assuming that the combustion gases do not change composition during the expansion stroke, calculate the indicated mean effective pressure (imep, p j ) of the cycle in bar. What has been the effect of dissociation on the power output of the engine? Assume that the index of compression K~ = 1.4, while that of expansion, lcE = 1.35. (See Chapter 10, Question 7).

[2487 K; 99.4 bar, 15.10 bar]

20 One method of reducing the maximum temperature in an engine is to run with a rich mixture. A spark ignition engine with a compression ratio of 10: 1, operating on the

Page 280: Advanced Thermodynamics for Engineers

264 Chemical equilibrium and dissociation

21

Otto cycle, runs on a rich mixture of octane and air, with an equivalence ratio of 1.2. The trapped conditions are 1 bar and 300 K, and the index of compression is 1.4. Calculate, taking into account dissociation of the carbon dioxide and water vapour, how much lower the maximum temperature is under this condition than when the engine was operated stoichiometrically. How has dissociation affected the products of combustion? What are the major disadvantages of operating in this mode? (See Chapter 10, Question 8.)

[3020 K, 3029 K]

A gas engine with a volumetric compression ratio of 10 : 1 is run on a weak mixture of methane (C&) and air, with @ = 0.9. If the initial temperature and pressure at the commencement of compression are 60°C and 1 bar respectively, calculate the maximum temperature and pressure reached during combustion at constant volume, taking into account dissociation of the carbon dioxide and water vapour, under the following assumptions:

(a) (b) that compression is isentropic.

Assume the ratio of specific heats, K, during the compression stroke is 1.4, and the heat of reaction at constant volume for methane is -8.023 x lo5 kJ/kmol C h . (See Chapter 10, Question 9.)

[2737 K; 71.57 bar]

that 10% of the heat released is lost during the combustion period, and

Page 281: Advanced Thermodynamics for Engineers

13 Effect of Dissociation on Combustion

Pa ra meters

Dissociation is an equilibrium process by which the products of combustion achieve the minimum Gibbs function for the mixture (see Chapter 12). The effect is to cause the products that would be obtained from complete combustion to break down partially into the original reactants, and other compounds or radicals. This can be depicted on a U-T diagram as shown in Fig 13.1.

- , .

,' . I .

I . , . I .

I .

Products

' ' Products (with dissociation)

(with complete combustion)

T, for issociation

T, for complete combustion 7

Temperature, T

Fig. 13.1 Internal energy-temperature diagram for combustion both with and without dissociation

It can be seen from Fig 13.1 that the effect of dissociation is to reduce the temperature of the products after combustion. This, in turn, reduces the amount of energy that can be drawn from the combustion process and reduces the work output of engines.

The other effect of dissociation is to form pollutants. This is particularly true in combustion in engines, when carbon monoxide (CO) can be formed even when the mixture is weak - that is, there is more than sufficient oxygen in the air to oxidise completely both the carbon and hydrogen in the fuel. If chemical kinetics are considered (see Chapter 14), i.e. account is taken of the finite rate by which equilibrium is achieved, then it is possible to show how some of the major pollutants are formed in the quantities measured at the exit from engines. Chemical kinetic effects and dissociation are the sole

Page 282: Advanced Thermodynamics for Engineers

266 Effect of dissociation on combustion parameters

causes of the production of NO, from fuels which do not contain nitrogen (through combining of the nitrogen and oxygen in the air), and they also increase the amount of carbon monoxide (CO) during weak and slightly rich combustion. Processes which include chemical kinetics are not equilibrium processes, but they attempt to reach equilibrium if there is sufficient time.

The effect of chemical, or rate, kinetics can be assessed by considering a simple example. Imagine a spark ignition engine in which the mixture is compressed prior to ignition by a spark. If the compression temperature is not too high it can be assumed that the reactants do not react before ignition. After ignition the reactants bum to form products at a high temperature, and these are initially compressed further (as the piston continues to rise to top dead centre [tdc], and the combustion process continues) before being expanded when the piston moves down and extracts work from the gases. As the pressure increases, the temperature of the gas also increases and the products tend to dissociate: they attempt to achieve an equilibrium composition but the speed of the engine is too rapid for this. The effect can be seen in Fig 13.2, which depicts the way in which the actual level of pollutants attempts to follow the equilibrium level, but lags the equilibrium values. This lag means that the maximum level of pollutant (e.g. NO,) formed does not achieve the maximum equilibrium value, but it also means that the rate of reduction of the pollutants is slower than that of the equilibrium species and the NO, is usually frozen at a level above the equilibrium value of the exhaust gas from the combustion chamber, boiler, gas turbine, diesel engine, etc.

._.. . .... ... .. . ... ,., %,

.(....._

Crankangle

Fig. 13.2 Production of pollutants in equilibrium and chemical kinetically controlled processes

Page 283: Advanced Thermodynamics for Engineers

The basic reactions 267

13.1 Calculation of combustion both with and without dissociation It is possible for the pressure, temperature and equilibrium composition of various mixtures to be calculated using computer programs, based on the principles outlined in Chapter 12. A program written at UMIST and based on the enthalpy coefficients given in Chapter 9 has been used to evaluate the results presented in this chapter. The program, which has a range of fuels programmed into it, can be run in dissociation and non-dissociation modes. A series of calculations was undertaken using octane (C8H18) and methane (CHJ as the fuels. The equivalence ratio was varied from a weak value of @ = 0.5 to a rich one of @ = 1.25. It was assumed that the initial pressure and temperature at entry to the combustion chamber were 1 bar and 300 K respectively, and that the fluid was compressed through a volumetric compression ratio of 12, with an index of compression (K = c,/c,) of 1.4. Combustion then took place at constant volume. This is essentially similar to the combustion that would take place in a reciprocating engine with constant volume combustion at tdc, which is an ‘Otto cycle’ in which the heat transfer process associated with the air standard cycle is replaced by a realistic combustion process. Combustion in this Otto cycle is adiabatic and occurs at constant volume: this is a constant internal energy process, as shown in Fig 13.1. The principles introduced here can be developed to study the effect of equilibrium on a realistic (finite rate of combustion) cycle.

The computer program was written in a manner that enabled it also to be used for constant pressure combustion, e.g. like that occurring in a gas turbine combustion chamber. In this case the enthalpy of both the reactants and products will be equal.

13.2 The basic reactions The stoichiometric combustion of methane is defined by the chemical equation

CH4 + 2(02 + 3.76N2) * CO, + 2H20 + 7.52N2 (13.1)

while that for octane is

CsH1, + 12.5(02 + 3.76N2) 8C02 + 9H,O + 47N, (13.2)

If the mixture is not stoichiometric then the equation for methane becomes

7.52 N2 (13.3)

if the mixture is weak, i.e. there is less oxygen than required for complete combustion of the fuel. @ is called the equivalence ratio and in this case @ s 1.0. For a rich mixture, @ 3 1.0, and if it is assumed that the hydrogen has preferential use of the oxygen, then the equation becomes

(: 1 @ 2

@ CH, + - ( 0 2 + 3.76Nz) 3 C02 + 2 - - 1 0 2 + 2H2O + -

7.52 CH4 + - (0, + 3.76N2) ( - c o , + ( ___ 4@; 4)co + 2H20 + - N, (13.4)

2

@ @ It will be assumed for the initial examples that the only dissociation is of the products,

carbon dioxide and water, which dissociate according to the equations

co, w co + io2 and

H20 w H, + i 0,

(13.5)

(13.6)

Page 284: Advanced Thermodynamics for Engineers

268 Effect of dissociation on combustion parameters

These equations (eqns (13.5) and (13.6)) have to be added into the basic equations to evaluate the chemical equation with dissociation. The program does this automatically and evaluates the mole fractions of the products. The approach adopted by the program is the same as that introduced in Chapter 12, when the products of combustion were defined by a set of simultaneous equations. Later in the chapter the dissociation of other compounds will be introduced to give a total of 11 species in the products. The method of solving for these species is outlined by Baruah (Chapter 14) in Horlock and Winterbone (1986).

13.3 The effect of dissociation on peak pressure

In general, dissociation will tend to decrease the pressure achieved during the combustion process (when it occurs in a closed system) because it reduces the temperature of the products. This reduction in pressure is always evident with stoichiometric and lean mixtures, although an increase in pressure over the equivalent situation without dissociation can occur in rich mixtures due to the increase in the amount of substance in the products. Figures 13.3 and 13.4 show that the peak pressure is reduced by about 8 bar at the stoichiometric air-fuel ratio (@ = 1) with both methane and octane. Dissociation also changes the equivalence ratio at which the peak pressure is achleved. If there is no dissociation then the peak pressure is always reached at the stoichiometric ratio (@ = 1). However, when dissociation occurs the equivalence ratio at which the peak pressure occurs is moved into the rich region (@ > 1). This is because dissociation tends to increase the amount of substance (np) in the products, compared with the non-dissociating case. The peak pressure with dissociation occurs at around @ = 1.1 for methane, and around @ = 1.25 for octane. It is interesting to note from the octane results (Fig 13.4) that the peak pressures achieved with and without dissociation are almost the same - except they occur at a different equivalence ratio.

n L

Rich . 80 a, L

a 20

0 0 5 O h 07 O X 0 9 1 1 1 1 2

Equivalence ratio, 4 Fig. 13.3 Variation of pressure with equivalence ratio for combustion of methane. Initial pressure:

1 bar; initial temperature: 300 K; compression ratio: 12; compression index: 1.4

13.4 The effect of dissociation on peak temperature

Figures 13.5 and 13.6 show the variation of the peak temperature, produced by an adiabatic combustion process, with equivalence ratio, both with and without dissociation.

Page 285: Advanced Thermodynamics for Engineers

The effect of dissociation on the composition of the products 269

I !

I y t I I

Equivalence ratio, 4, Fig. 13.4 Variation of pressure with equivalence ratio for combustion of octane. Initial pressure:

1 bar; initial temperature: 300 K; compression ratio: 12; compression index: 1.4

140 - 120 .’

n L m 100 P, . 80 aJ L 3 60 . m m a k

0 .

- - - - . !s”- -- ’ ,/I I

j/ i I .,-, I h

I I Meak Rich !

I 1 I

1

I 1

40.- - - - no dissociation ,

, 1

.. I 1/ 1 !

I

I t--.-- A

A- , _-i-- - dissociation

20.- , i I I i

Dissociation lowers the temperature for both fuels, and moves the point at which the maximum temperature is achieved. In the case of methane the maximum temperature reduces from 3192 K (at stoichiometric) to 3019 K (at about $J = 1.1 - rich), while for octane the values are 3306 K to 3096 K (at @ = 1.2). The effect of such a reduction in temperature is to lower the efficiency of an engine cycle.

+ m L 1500

E 1000 . 500 .

% F-”

13.5 The effect of dissociation on the composition of the products

The composition of the fuel will affect the composition of the products due to the different stoichiometric air- fuel ratios and the different carbon/hydrogen ratios. The stoichiometric air-fuel ratio (by weight) for methane is 17.16, whilst it is 15.05 for octane. The stoichiometric air-fuel ratio remains in the region 14 to 15 for most of the higher straight hydrocarbon fuels.

I I

, I I - ~ _ _ -c - __

~ .--~- I I

I I , I - - - no dissociation - dissociation !

I -

I I

Page 286: Advanced Thermodynamics for Engineers

270 Effect of dissociation on combustion parameters

- dissociation

0 5 0 6 0 7 0 8 0 9 1 I t 1 2 1 3

Equivalence ratio, 4 Fig. 13.6 Variation of temperature with equivalence ratio for combustion of octane. Initial pressure:

1 bar; initial temperature: 300 K, compression ratio: 12; compression index: 1.4

The carbon/hydrogen ratio varies from 0.25 to 0.44 for methane and octane respectively, and this affects the composition of the products, both with and without dissociation. The mole fractions of the products for the combustion of methane (taking account of dissociation) are shown in Fig 13.7. It can be seen that the largest mole fraction is that of water, which is more than twice that of carbon dioxide: this reflects the greater proportion of hydrogen atoms in the fuel. (Note: the mole fraction of

0. e 0

u a .I *,

t

0.20

0.15

0.10

0.05

0.00

I 1 I 9 I

I I H>O

0.50 0.60 0.70 0.80 0.90 1.00 1.10 1.20

Equivalence ratio, 4 Fig. 13.7 Variation of composition of products with equivalence ratio for combustion of methane.

Initial pressure: 1 bar; initial temperature: 300 K; compression ratio: 12; compression index: 1.4

nitrogen, N,, has been omitted from Figs 13.7, 13.8, 13.9 and 13.10 because it dominates the composition of the products. Its mole fraction remains around 0.79 in all cases.)

The mixture is weak at equivalence ratios of less than unity, and there is excess oxygen for complete combustion up to an equivalence ratio of unity. This can be seen in Figs 13.8

Page 287: Advanced Thermodynamics for Engineers

The effect offuel on composition of the products 27 1

0.10

& 0.08 I

c 0 P 0.06 0 Q

a 0.04 bk - f 0.02

0.00 0.50 0.60 0.70 0.80 0.90 1.00 1.10 1.20

Equivalence ratio, $ Fig. 13.8 Variation of carbon reactions, with and without dissociation, for combustion of methane

x

0.20

0.15

0.10

0.05

0.00

- dissociation

0.50 0.60 0.70 0.80 0.90 1.00 1.10 1.20

Equivalence ratio, 4 Fig. 13.9 Variation of hydrogen reactions, with and without dissociation, for combustion of

methane

and 13.9, where the composition of the products, both with and without dissociation, has been plotted. It can be seen from Fig 13.9, which includes the mole fraction of oxygen, that there is no oxygen in the products at equivalence ratios of greater than unity if there is no dissociation. However, when there is dissociation the quantity of oxygen in the products is increased, simply due to the reverse reaction. This reverse reaction also produces some unreacted hydrogen, and dissociation causes this to be produced even with weak mixtures. It should be noted that it was assumed that the hydrogen would be oxidised in preference to the carbon in the reactions without dissociation, and this explains the absence of hydrogen in those cases.

The carbon-related reactions are shown in Fig 13.8. When dissociation is neglected, the carbon is completely oxidised to carbon dioxide in the weak and stoichiometric mixtures, with no carbon monoxide being formed. After the mixture becomes rich, the level of carbon monoxide increases linearly with equivalence ratio. When dissociation is

Page 288: Advanced Thermodynamics for Engineers

272 Eflect of dissociation on combustion parameters

0. I F

K c e

0 0.10

b k

.* c, u m

2 0.05

P 0.0

0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3

Equivalence ratio, 4 Fig. 13.10 Variation of composition of products with equivalence ratio for combustion of octane.

Initial pressure: 1 bar; initial temperature: 300 K; compression ratio: 12; compression index: 1.4

considered, carbon monoxide is formed even in the weak mixture region, but surprisingly there is a lower level of carbon monoxide in the rich region than when there is no dissociation. The explanation of this is that the dissociation of the water releases some oxygen that is then taken up by the carbon. The production of carbon monoxide in lean mixtures is the reason for carbon monoxide in the untreated exhaust gas from internal combustion engines even when they are operating stoichiometrically or lean.

Considering the carbon dioxide, it can be seen that this peaks at an equivalence ratio of unity (@ = 1) without dissociation. However, it peaks at an equivalence ratio of around $I = 0.9 to 1.0 (i.e. weak) when there is dissociation, when the peak level of carbon dioxide is only about 75% of that without dissociation.

The mole fractions of products for the combustion of octane are shown in Fig 13.10. The trends are similar to those for the combustion of methane. The carbon dioxide fraction maximises on the lean side of stoichiometric, at @ of about 0.9. It is also noticeable that the mole fractions of water and carbon dioxide are much closer with octane than methane, which simply reflects the higher carbon/hydrogen ratio of octane.

13.6 The effect of fuel on composition of the products

Two different fuels were used in the previous analysis of the effect of dissociation on combustion products. Both are paraffinic hydrocarbons, with a generic structure of CnH2n+2: for methane the value of n = 1, while for octane n = 8. This means that the carbon/hydrogen ratio varies from 0.25 to 0.44. This has an effect on the amount of CO, produced for each unit of energy released (i.e. kg CO,/kJ, or more likely kg CO,/kWh). The effect of this ratio can be seen in Figs 13.7 and 13.10. The maximum mole fraction of CO, released with octane is almost lo%, whilst it peaks at around 7.5% with methane: the energy released per unit mass of mixture is almost the same for both fuels, as shown by the similarity of temperatures achieved (see Figs 13.5 and 13.6). Hence, if it is required to reduce the amount of carbon dioxide released into the atmosphere, it is better to bum fuels with a low carbon/hydrogen ratio. This explains, in part, why natural gas is a popular fuel at this time.

Page 289: Advanced Thermodynamics for Engineers

The formation of oxides of nitrogen 273

13.7 The formation of oxides of nitrogen

It was shown in Chapter 12 that when air is taken up to high temperatures, the nitrogen and oxygen will combine to form oxides of nitrogen, and in particular nitric oxide, NO. It was also stated that it is not enough simply to add the equation

N, + 0, w 2 N O (13.7)

but a chain of reactions including 11 species has to be introduced. This is necessary because the formation of atomic oxygen and nitrogen, and the radical OH play a significant part in the amount of NO produced in the equilibrium mixture.

These equations representing equilibrium of the 11 have been used to analyse the equilibrium constituents of the products of combustion for the cases considered above, i.e. the combustion of methane and octane in an engine operating on an ‘Otto’ cycle. The results are shown in Figs 13.11 and 13.12. It should be noted that the ordinate axis is in logarithmic form to enable all the important constituents to be shown on a single graph. The numerical results for carbon monoxide, carbon dioxide and water are similar to those in Figs 13.9 and 13.10; it is simply the change in scales which has changed the apparent shapes. These results will now be analysed.

1

2 0.1 c E 0

u Q

0)

.I 4-

& II

0.01 9

0.001

0.6 0.8 1 .0 1.2 Equivalence ratio, +

Fig. 13.11 Products of combustion at tdc for the combustion of methane in an Otto cycle

Page 290: Advanced Thermodynamics for Engineers

274 Effect of dissociation on combustion parameters

1

g 0.1 EI

k

0

u co

Q)

.H +

I

0.01 s

0.001

0.6 0.8 1.0 1.2 Equivalence ratio, 4

Fig. 13.12 Products of combustion at tdc for the combustion of octane in an Otto cycle

Considering Fig 13.11 first: this shows the equilibrium constituents of the mixture at tdc after adiabatic combustion of methane. The temperatures and pressures achieved by this combustion process are similar to those shown in Figs 13.3 and 13.5, although there will have been a slight reduction in both parameters because the dissociation processes all absorb energy. It can be seen that the maximum values of the mole fraction of both water (~0 .19 ) and carbon dioxide (~0.07) have not been significantly changed by the extra dissociation processes, and the CO, still peaks in the vicinity of q5 = 0.8 to 0.9. However, detailed examination shows that the levels of CO, and CO have been changed in the rich region, and in this case the mole fraction of CO exceeds that of CO, at q5 = 1.2. This is because more reactions are competing for the available oxygen, and the carbon is the loser.

Now considering the formation of nitric oxide: it was shown in Chapter 12 that nitrogen and oxygen will tend to combine at high temperatures to produce nitric oxide because the enthalpy of formation of nitric oxide is positive. This combination process will be referred to under the general heading of dissociation, although strictly speaking it is the opposite. This example is different from that shown in Chapter 12, when the variation in the amount of NO with equivalence ratio at constant pressure and temperature was depicted. In this case the peak pressure and temperature achieved in the combustion process are functions of the equivalence ratio, and the variations are shown in Figs 13.3 and 13.5. In fact, the

Page 291: Advanced Thermodynamics for Engineers

Effect of dissociation on combustion parameters 275

peak temperature increases from about 2200 K at Q, = 0.5, to 3000 K at Q, = 1.1; the peak pressure goes from about 87 bar to 122 bar over the same range of equivalence ratio. The increase in temperature will tend to increase dissociation effects in all cases, whereas an increase in pressure will tend to decrease the dissociation effects in cases where dissociation increases the total amount of substance (e.g. the CO, reaction); this was discussed in Chapter 12. Hence, the changes occurring in the conditions in the cycle tend to run counter to each other to some extent, although the effect of temperature dominates. The value of mole fraction of NO peaks at around Q, = 0.7 to 0.8, and obviously does not occur at the highest temperature. This is because the reaction which creates NO is dependent on the amounts of nitrogen (N,) and oxygen (0,) present in the mixture, as well as the temperature of the products. At low equivalence ratios there is an abundance of components to react but the temperature is low, whilst at high temperatures the driving force to react is high but there is a deficiency of oxygen. In fact, the nitrogen has to compete with the carbon monoxide and hydrogen for any oxygen that becomes available, and it gains most of the oxygen from the carbon reactions. The maximum level of NO attained is a mole fraction of about 0.13, and this is equivalent to about 13 000 ppm (parts per million [by mass]); this level is typical of that which would be achieved at equilibrium conditions in an engine.

Before considering the effect of the rate of reactions on the actual levels of NO achieved, it is worthwhile seeing the effect of fuel on the level of exhaust constituents. The additional reactions do not have a major effect on the exhaust composition with octane as a fuel, and the CO, level is still significantly higher than with methane, as would be expected. However, the general level of NO is higher with octane than methane because of the slightly higher temperatures achieved with this fuel, and the larger proportion of CO, which can be ‘robbed’ for oxygen. The maximum levels of NO again occur around Q, = 0.7 to 0.8, and result in about 14 000 ppm.

The levels of NO in the exhaust gas of an engine are significantly lower than those calculated above for a number of reasons. First, the reactions occurring in the cylinder are not instantaneous, but are restricted by the rate at which reacting molecules meet; this is known as chemical kinetics or rate kinetics, and is the subject of Chapter 14. The effect of these kinetics is that the instantaneous level of NO in the cylinder does not achieve the equilibrium level, and lags it when it is increasing as shown in Fig 13.2. When the equilibrium concentration is decreasing, the actual level of NO still cannot maintain pace with it, and again lags the level while it decreases down the expansion stroke of the engine, as shown qualitatively in Fig 13.2; in fact these reactions tend to ‘freeze’ at around 2000K, and this level then dominates the exhaust value. Secondly, the combustion process does not take place instantaneously at tdc as shown here, but is spread over a significant period of the cycle and the pressures and temperatures are lower than those attained in the ‘Otto’ cycle: this will reduce the peak equilibrium level of NO.

It is shown in Figs 13.11 and 13.12 that the peak level of NO at equilibrium occurs at around Q, = 0.7 to 0.8, but experience shows that the maximum values of NO in engine exhaust systems (around 2000 to 5000 ppm) occur at about @ = 0.9. This can be explained by considering the effect of rate kinetics. It will be shown in Chapter 14 that the rate at which a reaction occurs is exponentially related to temperature. Hence, while the equilibrium level of NO is a maximum at g5 = 0.7 to 0.8, the lower temperature at this equivalence ratio limits the amount of NO produced by the reaction. The peak value of NO occurs at Q, = 0.9 because the driving force (the equilibrium concentration) and the rate of reaction combine at this equivalence ratio to maximise production of this pollutant.

Page 292: Advanced Thermodynamics for Engineers

14 Chemical Kinetics

14.1 Introduction

Up until now it has been assumed that chemical reactions take place very rapidly, and that the equilibrium conditions are reached instantaneously. While most combustion processes are extremely fast, often their speed is not such that combustion can be considered to be instantaneous, compared with the physical processes surrounding them. For example, the combustion in a reciprocating engine running at 6000 rev/min has to be completed in about 5 ms if the engine is to be efficient. Likewise, combustion in the combustion chamber of a gas turbine has to be rapid enough to be completed before the gas leaves the chamber. Such short times mean that it is not possible for all the gases in the combustion chamber to achieve equilibrium - they will be governed by the chemical kinetics of the reactions.

Chemical kinetics plays a major role in the formation of pollutants from combustion processes. For example, oxygen and nitrogen will coexist in a stable state at atmospheric conditions, and the level of oxides of nitrogen (NO,) will be negligible. However, if the oxygen and nitrogen are involved in a combustion process then they will join together at the high temperature to form NO, which might well be frozen into the products as the temperature drops. This NO, is a pollutant which is limited by legislation because of its initant effects. NO, is formed in all combustion processes, including boilers, gas turbines, diesel and petrol engines; it can be removed in some cases by the use of catalytic converters.

It was shown in Chapters 12 and 13 that significant dissociation of the normal products of combustion, carbon dioxide and water can occur at high temperature. The values shown were based on the equilibrium amounts of the substances, and would only be achieved after infinite time; however, the rates at which chemical reactions occur are usually fast and hence some reactions get close to equilibrium even in the short time the gases are in the combustion zone. An analysis of the kinetics of reactions will now be presented.

14.2 Reaction rates

Reaction rates are governed by the movement and breakdown of the atoms or molecules in the gas mixture: reactions will occur if the participating ‘particles’ collide. The number of

Page 293: Advanced Thermodynamics for Engineers

Reaction rates 277

collisions occurring in a mixture will be closely related to the number densities (number per unit volume) of the 'particles'. The number density can be defined by the molar concentration, c , which is the amount of substance per unit volume. This is obviously a measure of the number of particles per unit volume since each amount of substance is proportional to the number of molecules; this is illustrated in Fig 14.1. The molar concentration will be denoted by enclosing the reactant or product symbol in [ 1.

. . . 0 . .

. 0 .

. O . . . . 1: 'd.1

Fig. 14.1 Diagrammatic representation of molar concentration. Molar concentration in (a) is approximately half that in (b)

Reactions occur when two, or more, reactants are capable of reacting. Many simple chemical reactions are second order, e.g.

kf A + B - C + D

k b (14.1) kf

C02 + H2-H20 + CO k b

where the first reaction is a general one and the second is an example based on the water gas reaction (chosen because the same number of reactants and products exist on both sides).

These reactions can be written as

(14.2)

where A , , A,, .. . etc are the elements or compounds involved in the reaction, and Y , , v2 , . . . etc are the respective stoichiometric coefficients. Equation (14.2) can be further generalised to

(14.3)

where q = total number of species being considered, v' represents the stoichiometric coefficients of the reactants, and Y" that of the products.

In this case q has been taken as the same on both sides of the equation because it is assumed that the same species can exist on both sides. If a species does not exist on one of the sides it is represented by a stoichiometric coefficient of zero (Y = 0). For example,

(14.4) CO, + H, wH,O + CO

Page 294: Advanced Thermodynamics for Engineers

278 Chemical kinetics

can be represented by

4 b 4 v:,~i-.C V;A, i = 1 k b i = l

(14.5)

where q = 4 vij = stoichiometric coefficient of i in reaction j A , = C 0 2 v \ = l v ’ ~ = O

A, = H2 v ; = 1 v ; = o

A,= CO vk=O v:=1 A,=H,O v;=O ~ l ; = l

The law of mass action, which is derived from the kinetic theory of gases, states that the rate of formation, or depletion, of a species is proportional to the product of molar concentrations of the reactants, each raised to the power of its stoichiometric coefficient. Hence the rate of formation of species j , for an elementary reaction, is

Rj a [Al]”J[A2]’*J [A,]”

or n

Rj=n[Ai]” f o r ( j = 1, ... s) i = 1

and s = total number of simultaneous reactions. Equation (14.6) may be written as

n

R j = k , n [ A J y J f o r ( j = 1, ... s) i = 1

(14.6)

(14.7)

where kj is the rate constantfor the reaction. In eqns (14.1) to (14.5) the reaction was described as having a forward and backward direction, with rates k, and kb. The reaction shown in eqn (14.7) can be described in this way as

n

R, = bj [A,]’” for ( j = 1, . . . s) (14.8) i = 1

and n

Rbj = kbj n [A,]” for ( j = 1 , . . . S) (14.9)

The rate of change with time of species i is proportional to the change of the stoichiometric coefficients of Ai in the reaction equation (the system is effectively a first order one attempting to achieve the equilibrium state). Thus, for the forward direction

i = 1

(14.10)

Page 295: Advanced Thermodynamics for Engineers

Chemical kinetics of NO 279

and, for the backward direction

Hence the net rate of formation of Ai is

[A,]'' - kbj n [Ai]" d [Ail j

i = 1

Consider the following rate controlled reaction equation:

kf vaA + vbB-vcC + vdD

k b

The net rate of generation of species C is given by

-- d[C1 - [kf[A]'"[B]" - kb[c]''[DP] dt

(14.11)

(14.12)

(14.13)

( 14.14)

Using the notation [A]/[Al, = a , [Bl/[B], = /3, [Cl/[Cl, = y and [Dl/[Dl, = 6, where the suffix e represents equilibrium concentrations, gives

(14.15) d [ C ] /d t = k, a "0/3"b[A ] 2 [ B 12 - kbyVc 6 vd [ C 1,'. [ D ] 2 At equilibrium

k,[A]:[B]? = k,[C],'.[D],'d= Rj

Therefore the net rate is

(14.16)

d[C]/dt = Rj[av~/3vb - y"C6vd1 (14.17)

14.3 Rate constant for reaction, k

The rate constant for the reaction, k, is related to the ability of atoms or ions to combine. In combustion engineering this will usually occur when two or more particles collide. Obviously the collision rate is a function of the number of particles per unit volume, and their velocity of movement, i.e. their concentration and temperature. Most chemical reactions take place between two or three constituents because the probability of more than three particles colliding simultaneously is too small. It has been found experimentally that most reactions obey a law like that shown in Fig 14.2.

This means that the rate constant for the reaction can be defined by an equation of the form

(14.18)

This equation is called the Arrhenius equation. The factor A is called the pre- exponential factor, or the frequency factor, and is dependent on the rate at which collisions with the required molecular orientation occur. A sometimes contains a temperature term, indicating that the number of collisions is related to the temperature in those cases. The term E in the exponent is referred to as the activation energy. The values of A and E are dependent on the reactions being considered, and some values are

k = Ae - E / R T

Page 296: Advanced Thermodynamics for Engineers

280 Chemical kinetics

11 T

Fig. 14.2 Relationship between reaction rate and temperature

introduced below. The significance of the activation energy, E , is shown in Fig 14.3, where it can be seen to be the energy required to ionise a particular molecule and make it receptive to reacting. It appears in the exponential term because not all activated molecules will find conditions favourable for reaction.

Activation

Enthalpy (or internal energy) of reaction, Q, (or Q, )

Progress of reaction

Fig. 14.3 Schematic interpretation of activation energy

14.4 Chemical kinetics of NO

The chemical kinetics for the formation of NO are relatively well understood, and will be developed here. The chemical kinetics for other pollutants can be derived in a similar way if the necessary reaction rates are available, although it should be recognised that most other pollutants are produced by reactions between the oxygen in the air and a constituent of the fuel (e.g. carbon or sulfur). The formation of NO in combustion processes can occur from two sources: thermal NO and prompt NO. Thermal NO is formed by the combination of the oxygen and nitrogen in the air, and will be produced even if there is no nitrogen in the fuel itself. This section will restrict itself to considering thermal NO. Prompt NO is

Page 297: Advanced Thermodynamics for Engineers

Chemical kinetics of NO 281

thought to be formed in the flame as a result of the combination of the nitrogen in the fuel with the oxygen in the air. The amount of nitrogen in most conventional hydrocarbon fuels is usually very low.

The governing equations for the mechanism of NO formation are (Lavoie et al. 1970)

(1) N + N O e N , + O , k,, = 3.1 x 10" x e(-'60/T) m3/kmol s ( 1 4.1 9)

(2) N+O,-NO+O, k,, = 6.4 x lo6 x T x m3/kmol s (14.20)

(3) N+OH-NO+H, k,, = 4.2 x iolo m3/kmoi s (14.21)

(4) (14.22)

( 5 ) 0 + N , O e N , + 0,, k f5 - - 3.2 1012 ,(-189W/T) m3/kmol s (14.23)

(6) 0 + N,O NO + NO, k,, = kf5 (14.24)

(7) (14.25)

In these equations the rate constants (kf , ) are all in m3/kmol s. M is a third body which may be involved in the reactions, but is assumed to be unchanged by the reactions. M can be assumed to be N,. These equations can be applied to the zone containing 'burned' products, which exists after the passage of the flame through the unburned mixture. It will be assumed that H and OH, and 0 and 0, are in equilibrium with each other; these values can be calculated by the methods described in Chapter 12.

H + N,O=N, + OH, k f4 - - 3.0 x 10'0 x e(-535'J/T) m3/kmol s

N,O + M e N , + 0 + M, k , = lo', x e(-305W/T) m3/kmo1 s

14.4.1

The rate of formation of nitric oxide can be derived in the following manner. Consider reaction j : let k, be the forward reaction rate, k, the backward rate, and R, the 'one way' equilibrium rate. Also let

RATE EQUATIONS FOR NITRIC OXIDE

[NOI/[NOIe= a , [NI/[NIe = B, [N,OI/[N,OIe= Y where suffix e denotes equilibrium values. Then the following expressions are obtained for the burned gas, or product, volume.

Expression for nitric oxide, NO From eqn (14.19) the net rate is

-kfl[Nl [NO1 + k,I[N21 [OI = -aBkf,[Nle[NOle + kbl[N2le[Ole

But

k,,[Nle[NOle= kbl[N2le[Ole= R,,

so the net rate for eqn (14.19) becomes -aBR, + R,. Using a similar procedure for eqns (14.20), (14.21) and (14.24) involving NO gives the following expression, which also allows for the change in volume over a time step:

1 d _ _ ([NOIV) = -a(BR, + R, + R, + 2aR6) + R, + B(R, + R,) + 2yR6 V dt

(14.26)

where V is the volume of the products zone.

Page 298: Advanced Thermodynamics for Engineers

282 Chemical kinetics

Expression for atomic nitrogen, N Equations (14.19), (14.20) and (14.21), which all involve N, give

1 d -- ([NIV) = -B(aR, + R, + R3) + R, + a(R, + R3) V dt

Expression for nitrous oxide N,O Equations (14.22) to (14.25) all involve N20 and can be combined to give

1 d - - ([N,O]V) = -y(R4 + R, + R, + R,) + V dt

+ R, + a'% + R,

(14.27)

(14.28)

A finite time is required for the reactions to reach their equilibrium values; this is called the relaxation time. It has been found (Lavoie et al. 1970) that the relaxation times of reactions (14.27) and (14.28) are several orders of magnitude shorter than those of reaction (14.26), and hence it can be assumed that the [N] and [N20] values are at steady state, which means that the right-hand sides of eqns (14.27) and (14.28) are zero. Then, from eqn (14.27)

Rl + a(R2 + R,) ( a K + R2 + R3)

B =

and from eqn (14.28)

R4 + R5 + a2R6 + R,

(% + R5 + R6 + R7) Y =

These values can be substituted into eqn (14.26) to give

(14.29)

(14.30)

This is the rate equation for NO that is solved in computer programs to evaluate the level of NO in the products of combustion. It should be noted that the variation in the molar concentration is a first-order differential equation in time and relates the rate of change of [NO] to the instantaneous ratio of the actual concentration of NO to the equilibrium value (i.e. a) . When the actual level of [NO] is at the equilibrium level then a = 1 and the rate of change of [NO] = 0.

14.4.2

Heywood et al. (1971) derived the initial rate of formation of NO in the following way. It can be shown that eqn (14.3 1) is the dominant equation for the initial formation of NO. When the nitric oxide (NO) starts to form, the value of a = 0, and eqn (14.31) can be written

INITIAL RATE OF FORMATION OF NO

(14.32)

Page 299: Advanced Thermodynamics for Engineers

Chemical kinetics of NO 283

Heywood (1988) stated that, using a three equation set for the formation of NO, the value of the rate constant for the reaction in the forward direction is

k,, = 7.6 x 1013 x e(-38000/T) cm3/mol s (14.34)

It is necessary to note two factors about eqn (14.34): the rate constant for the reaction is in cm3/mols, and the exponential term is significantly different from that in eqn (14.19). Hence

Heywood (1988) also showed that

where

Substituting this value into eqn (14.34) gives

[02]~/2[N2], cm3/mols d[NO1 - e(-69090/T)

dt T1/2

(14.35)

(14.36)

(14.37)

The values of [O,], and [N2],, which should be in mol/cm3, can be obtained from an equilibrium analysis of the mixture. The value of [O,], can be calculated from the perfect gas law, because

Likewise

(14.38)

(14.39)

Calculating the initial rate of change of NO with time gives the results shown in Fig 14.4. The conditions used to evaluate Fig 14.4 were: pressure, 15 bar; fuel, octane, which are based on those in Heywood (1988). Using these conditions the equilibrium concentrations of all the reactants were calculated using the techniques described in Chapter 12. This gave the following mole fractions for the condition T = 2000 K; @ = 0.6:

x0 =4.7725 x lo-’; xNZ = 7.5319 x lo-’; xoz = 7.7957 x lo-,

The rate kinetic values, which depict the rate of change of mole fraction of NO, were then calculated in the following way. Equation (14.37) is based on cm3, mol and seconds rather than SI units of m3, kmol and seconds and this must be taken into account.

Page 300: Advanced Thermodynamics for Engineers

284 Chemical kinetics

7 f

7- / -

2000 7100 2200 2300 2400 2500 2600 2700 2800

Temperature / (K) Fig. 14.4 Initial rate of formation of nitric oxide for combustion of octane at a pressure of 15 bar

The molar concentration is defined as

xun xu

V vm

[MI = - = - , where M is a general substance

%T 8.3143 x 2000 Vm=-= x lo6 = 11086 cm3/mol

P is io5

x, 4.7725 x lo-’ [O], = - - - = 4.3051 x moi/cm3

11086 1 1086

-XN> 7.5319 x lo-’ m 1 e = - - - = 6.7941 x lo-’ mol/cm3

1 1086 11086

(14.40)

(14.41)

$ 7.7957 x lo-* [Q], = - = - = 7.0320 x mol/cm’

11086 11086

Page 301: Advanced Thermodynamics for Engineers

Chemical kinetics of NO 285

Using a combination of eqns (14.34) and (14.35) gives

dt

x 4.3051 x lo-’ x 6.7941 x (-38000/2000) = 15.2 x 1013 x e

= 2.491 x mol/cm3 s

A similar value can be obtained from eqn (14.37), which gives

(14.42)

=--- x e ( - 6 9 0 9 0 ~ 2 ~ ) x (7.0320 x 10-6)’/2 x 6.7941 x

= 2.4022 x mol/cm3 s (14.43)

These values can be converted from molar concentrations to more useful parameters and, in this case, they will be depicted as rate of formation of mole fraction. Rearranging eqn (14.40) gives

(14.44)

20001/2

xM = [Mlv,, where M is a general substance

Hence, the rate of production of NO in terms of mole fraction is given by

For this condition

d(XNO) - = 11086 x 2.491 x lo-’ = 2.762 x s-’ dt

(14.45)

(14.46)

The curves in Fig 14.4 were calculated in this manner. It can be seen that either eqns (14.34) and (14.35), or eqn (14.37) give similar results. This indicates that the dissociation equation, eqn (14.36), for oxygen is similar to that used in the program for calculating the values of mole fraction of atomic oxygen.

Other kinetics-controlled pollutants

Carbon monoxide (CO) is formed in processes in which a hydrocarbon is burned in the presence of oxygen. If the mixture is rich (Le. there is more fuel than oxygen available to oxidise it) then there is bound to be some carbon monoxide formed. However, even if the mixture is lean there will be some carbon monoxide due to the dissociation of the carbon dioxide; this was discussed in Chapters 12 and 13. The amount of carbon monoxide reduces as the mixture gets leaner, until misfire occurs in the combustion device.

Sulfur dioxide (SO,) is formed by oxidation of the sulfur in the fuel. Carbon particulates are formed during diffusion combustion because the oxygen in the

air is not able to reach all the carbon particles formed during the pyrolysis process while they are at a suitable condition to bum.

Page 302: Advanced Thermodynamics for Engineers

286 Chemical kinetics

14.5 The effect of pollutants formed through chemical kinetics

14.5.1 PHOTOCHEMICAL SMOG

Photochemical smogs are formed by the action of sunlight on oxides of nitrogen, and the subsequent reactions with hydrocarbons. Photochemical smogs were first identified in Los Angeles in the mid-l940s, and brought about the stringent legislation introduced in California to control the emissions from automobiles. These smogs act as bronchial irritants and can also irritate the eyes.

Among the pollutants involved in photochemical smogs are ozone, nitrogen dioxide and peroxyacyl nitrate (PAN). The nitrogen dioxide, and other oxides of nitrogen, are primary pollutants produced by dissociation in combustion reactions, and both ‘prompt’ and ‘thermal’ NO, can be involved in the reactions. Ozone and PAN are secondary pollutants produced by the action of sunlight on the primary pollutants and the atmosphere. Glassman (1 986) described in some detail the chemical chain that results in photochemical smog.

Oxides of nitrogen can be measured using non-dispersive infra-red (NDIR) equipment and the values are often quoted in parts per million (ppm), g/m3 (at normal temperature and pressure), g/kWh, or g/mile (if applied to a vehicle engine).

14.5.2 SULFUR DIOXIDE (SO,) EMISSIONS

Sulfur emissions can only occur if there is sulfur in the fuel. The levels of sulfur in diesel fuel and petrol are being reduced at the refinery in the developed world, and the sulfur levels are often around 0.5% or lower. SO, emissions from vehicles are not a major problem. In the developing world, the levels of sulfur might be significantly higher and then the SO, emissions from vehicles become a more serious problem. A further advantage of reducing the sulfur content of diesel fuels is that this reduces the synergy between the sulfur and the particulates, and reduces the quantity of particulates produced. The levels of sulfur dioxide produced in some power plant can be extremely high. Diesel engines running on ‘heavy fuel’, which has a high percentage of sulfur, will produce significant quantities of sulfur dioxide. Power stations running on heavy fuels or some coals will produce large quantities of sulfur fumes, which smell noxious, cause corrosion and form acid rain. Many coal burning power stations are now being fitted with flue gas desulfurisation, a process by which the sulfur dioxide reacts with limestone to form gypsum. These processes are suitable for stationary plant, but do consume copious amounts of limestone.

Sulfur in fuel has been a problem for a long time and caused high levels of acid rain in UK until recently. The move away from coal to gas for electricity generation, and the imposition of smokeless zones in urban areas have reduced the problem to a more acceptable level. Sulfur can be removed from crude oil by catalytic hydrodesulfurisation, but is often left in the residual oils which are used for marine propulsion, power generation, etc. The use of orimulsion, a fuel obtained by water extraction of low-grade oils, also brings worries of high sulfur fuels.

Sulfur in the fuel tends to be oxidised into SO, during the combustion process. This can then form sulfurous (H2S03) and sulfuric (H,SO,) acids, which have corrosive effects on exhaust stacks, chimneys, etc., before being released into the surrounding atmosphere to form acid rain. It is possible to use flue gas desulfurisation to reduce the levels of SO, emitted to the atmosphere, but such systems are expensive to install and run. It has been

Page 303: Advanced Thermodynamics for Engineers

The effect of pollutants formed through chemical kinetics 287

stated that if all the SO, emitted from the UK power stations were converted to sulfuric acid the supply of this acid would exceed the UK demand.

14.5.3 PARTICULATE EMISSIONS

These are generated mainly by devices which rely on diffusion burning (see Chapter 15), in which a rich zone can produce carbon particles prior to subsequent oxidation. Hence, particulate emissions tend to come from diesel engines, gas turbines and boiler plant. The carbon formed during combustion can be an asset in a boiler because it increases the radiant heat transfer between the flame and the tubes of the boiler; however, it is necessary to ensure that the soot emissions from the stack are not excessive. The carbon formed in diesel engine and gas turbine combustion is not beneficial and it increases the heat transfer to the components, increasing the need to provide cooling. The words particulates and soot are often used synonymously, but there is a difference in nature between these emissions. Dry soot is usually taken to mean the carbon that is collected on a filter paper in the exhaust of an engine. The unit of measurement of soot is usually the Bosch Smoke Number, which is assessed by the ‘reflectance’ of a filter paper on which the soot has been collected. Particulates contain more than simply the dry soot; they are soot particles on which other compounds (often polycyclic aromatic hydrocarbons [PAH]) have condensed. The PAH compounds have a tendency to be carcinogenic. The level of particulates is strongly affected by the amount of sulfur in the fuel, and increases with sulfur content. Particulates are measured by trapping the particles on glass-fibre filter papers placed in a dilution tunnel, and then weighing the quantity; they might be quoted in g/kW h.

14.5.4 GREENHOUSE EFFECT

The greenhouse effect is the name given to the tendency for the carbon dioxide in the atmosphere to permit the passage of visible (short wavelength) light, while absorbing the long wavelength infra-red transmission from objects on Earth. The effect of this is to cause the temperature of the Earth to increase. The evidence for global warming is not conclusive because the temperature of the Earth has never been constant, as shown by the Ice Ages of the past. There have been significant variations in temperature through the last few centuries, with warm summers at times and the freezing over of the Thames at others. It will take some time to prove conclusively whether global warming is occurring. What is indisputable is that mankind has released a large amount of carbon from its repository in hydrocarbons back into the atmosphere as CO,. The most convenient form of energy available to man is that found as gas, oil or coal. These are all hydrocarbon fuels with the quantity of hydrogen decreasing as the fuel becomes heavier. The combustion of any of these fuels will produce carbon dioxide.

Obviously, the policy that should be adopted is for the amount of CO,/kW or CO,/mile to be reduced. This can be achieved by using more efficient engines, or changing the fuel to, say, hydrogen. The most efficient power station, using conventional fuels, in terms of CO,/kW is the combined cycle gas turbine plant (CCGT) running on methane. Such plant can achieve thermal efficiencies higher than 50% and claims of efficiencies as high as 60% have been made. Large marine diesel engines also achieve thermal efficiencies of greater than 50%, and other smaller diesel engines can achieve around 50% thermal efficiencies when operating at close to full-load. It is presaged that the inter-cooled

Page 304: Advanced Thermodynamics for Engineers

288 Chemical kinetics

regenerated gas turbine will also have efficiencies of this order, and it is claimed, will compete with turbo-charged diesel engines down to the 5 M W size range.

The hydrogen powered engine produces no carbon dioxide, but does form NO,. However, the major problem with hydrogen is its production and storage, particularly for mobile applications. California has now introduced the requirement for the zero emissions vehicle (ZEV) which must produce no emissions. At present the only way to achieve this is by an electric vehicle. It has been said that a ZEV is an electric car running in California on electricity produced in Arizona! This is the major problem that engineers have to explain to politicians and legislators: the Second Law states that you cannot get something for nothing, or break-even. Many ideas simply move the source of pollution to somewhere else.

14.5.5 CLEAN-UP METHODS

The clean-up methods to be adopted depend upon the pollutant and the application. As stated above, the exhaust of a power station can be cleaned up using a desulfurisation plant to remove the sulfur compounds. It is also possible to remove the grit from the power station boilers by cyclone and electro-static precipitators. These plants tend to be large and would be inappropriate for a vehicle, although investigations to adapt these principles to diesel engines are continuing.

All petrol-driven cars being sold in the USA, Japan and Europe are fitted with catalytic converters to clean up the gaseous emissions. This currently requires that the engine is operated at stoichiometric mixture strength so that there is sufficient oxygen to oxidise the unburned hydrocarbons (uHCs) while enabling the carbon monoxide (CO) and NO, to be reduced. This means that the engine has to be operated under closed loop control of the air-fuel ratio, and this is achieved by fitting a h sensor (which senses the fuel-air equivalence ratio) in the exhaust system. The error signal from the sensor is fed back to the fuel injection system to change the mixture strength. The need to operate the engine at stoichimetric conditions at all times has a detrimental effect on the fuel consumption, and investigations into lean operation (reducing) catalysts are being undertaken.

At present it is not possible to use catalytic converters on diesel engines because they always operate in the lean bum regime. This causes a problem because there is no mechanism for removing the NO, produced in the diesel engine combustion process, and it has to be controlled in the cylinder itself. Lean bum catalysts, based on zeolites, are currently under research and will have a major impact on engine operation - both spark- ignition and diesel.

Another pollutant from diesel engines is the particulates. These are a significant problem and attempts have been made to control them by traps. Cyclone traps are being investigated, as well as gauze-based devices. A major problem is that after a short time the trap becomes full and the carbon has to be removed. Attempts have been made to achieve this by burning the particulates, either by self-heating or external heating. At present this area is still being investigated.

14.6 Other methods of producing power from hydrocarbon fuels

All of the devices discussed above, i.e. reciprocating engines, gas turbines, steam turbines, etc., produce power from hydrocarbon fuels by using the energy released from the combustion process to heat up a working fluid to be used in a ‘heat engine’. Hence, all of

Page 305: Advanced Thermodynamics for Engineers

Problems 289

these devices are limited by the Second Law efficiency of a heat engine, which is itself constrained by the maximum and minimum temperatures of the working cycle. It was shown in Chapters 2 and 4 that the maximum work obtainable from the combustion of a hydrocarbon fuel was equal to the change in Gibbs energy of the fuel as it was trans- formed from reactants to products: if there was some way of releasing all of this energy to produce power then a Second Law efficiency of 100% would be achievable. A device which can perform this conversion is known as a fuel cell; this obeys the laws of thermodynamics but, because it is not a heat engine, it is not constrained by the Second Law. Fuel cells are discussed in Chapter 17.

14.7 Concluding remarks

'This is the first time non-equilibrium thermodynamics has been introduced. It has been shown that, while most thermodynamics processes are extremely rapid, it is not always possible for them to reach equilibrium in the time available.

The effect of rate kinetics on the combustion reactions is that some of the reactions do not achieve equilibrium and this can be a major contributor to pollutants. The rate equations for nitric oxide have been developed.

Finally, the pollutants caused by combustion have been introduced, and their effects have been described.

PROBLEMS 1 A reaction in which the pre-exponential term is independent of temperature is found to

be a hundred times faster at 200°C than it is at 25°C. Calculate the activation energy of the reaction. What will be the reaction rate at lOOO"C?

[30 840 kJ/kmol; 13 8141

2 A chemical reaction is found to be 15 times faster at 100°C than at 25°C. Measure- ments show that the pre-exponential term contains temperature to the power of 0.7. Calculate the activation energy of the reaction. What will be the reaction rate at 700"C?

[31 433 kJ/kmol; 15 2041

3 The rate of formation of nitric oxide (NO) is controlled by the three reversible chemical reactions:

k,' 0 + N,-NO + N

ki- . + lcz

N+C&-NO+O k i

&+ N + OH-NO + H

k;

Page 306: Advanced Thermodynamics for Engineers

290 Chemical kinetics

Use the steady state approximation for the nitrogen atom concentration and the assumption of partial equilibrium for the reactions governing the concentrations of 0, O,, H and OH, show that

6 R + a

a R + 1 B = -

where B = [N]/[N],, 6 = [N2]/[N2],, a = [NO]/[NO],, R = R,/(R, + R3), and Rj is the equilibrium reaction rate of reaction j , and [ ] denotes the molar concentration, and [ 1, is the equilibrium molar concentration. Derive an expression for d[NO]/dt in terms of R,, B, a and 6.

At a particular stage in the formation of nitric oxide the values of R and a are 0.26 and 0.1 respectively. Why is 6 = 1 likely to be a good approximation in this case? What is the error if the rate of formation of NO is evaluated from the larger approximation d[NO]/dt = 2R,, rather than the equation derived in this question?

[3.64%]

4 The rate of change of mole concentration of constituent A in a chemical reaction is expressed as

-- - -k[A]" &41 dt

While mole concentration is the dominant property in the reaction it is much more usual for engineers to deal in mole fractions of the constituents. Show that the rate of change of mole fraction of constituent A is given by

where p = density. Also show how the rate of change of mole fraction is affected by pressure.

[dxA/dt a p"-']

Page 307: Advanced Thermodynamics for Engineers

15 Combustion and flames

15.1 Introduction

Combustion is the mechanism by which the chemical (bond) energy in a ‘fuel’ can be converted into thermal energy, and possibly mechanical, power. Most combustion processes require at least two components in the reactants - usually a fuel and an oxidant. The chemical bonds of these reactants are rearranged to produce other compounds referred to as products. The reaction takes place in a flume. There are three parameters which have a strong influence on combustion: temperature, turbulence and time. In designing combustion systems attention must be paid to optimising these parameters to ensure that the desired results are achieved. In reciprocating engines the time available for combustion is limited by the operating cycle of the engine, and it is often necessary to increase the turbulence to counterbalance this effect. In furnaces the time available for combustion can be increased by lengthening the path taken by the burning gases as they traverse the chamber.

There are two basically different types of flame: premixed and diffusion. An example of premixed flames occurs in conventional spark-ignition (petrol, natural gas, hydrogen) engines. In these engines the fuel and air are mixed (often homogeneously) during the admission process to the engine, either by a carburettor or low pressure inlet manifold fuel injection system. Ignition is initiated by means of a spark, which ignites a small volume of the charge in the vicinity of the spark plug; this burning region then spreads through the remaining charge as a flame front. This type of combustion mechanism can be termed flume traverses charge (FTC), and once combustion has commenced it is very difficult to influence its progress. The diffusion flame occurs in situations of heterogeneous mixing of the fuel and air, when fuel-rich and fuel-lean regions of mixture exist at various places in the combustion chamber. In this case the progress of combustion is controlled by the ability of the fuel and air to mix to form a combustible mixture - it is controlled by the diffusion of the fuel and air. An example of this type of combustion is met in the diesel engine, where the fuel is injected into the combustion chamber late in the compression stroke. The momentum of the fuel jet entrains air into itself, and at a suitable temperature and pressure part of the mixture spontaneously ignites. A number of ignition sites may exist in this type of engine, and the fuel-air mixture then bums as the local mixture strength approaches a stoichiometric value. This type of combustion is controlled by the mixing (or diffusion) processes of the fuel and air. Other examples of diffusion combustion are gas turbine combustion chambers, and boilers and furnaces.

Page 308: Advanced Thermodynamics for Engineers

292 Combustion andflames

These different combustion mechanisms have an effect on how the energy output of the combustion process can be controlled. In the homogeneous, premixed, combustion process the range of air-fuel ratios over which combustion will occur is limited by the flammabil- ity of the mixture. In old petrol engines the mixture strength varied little, and was close to stoichiometric at all operating conditions (the mixture was often enriched at high load to reduce combustion temperatures). In modem engines, attempts are made to run the engines with weaker mixtures (Le. lean-burn). At present, all spark-ignition engines operate with a throttle to limit the air flow at low loads. This is known as quantitative governing of the engine power because the quantity of charge entering the engine is controlled, and this controls the power output. Throttling the engine reduces the pressure in the inlet manifold, and effectively the engine has to pump air from the inlet manifold to the exhaust system, giving a negative pumping work of up to 1 bar mep. The advantage of lean-bum engines is that the load control can be partly by making the mixture leaner, and partly by throttling. Honda have an engine which will run as lean as 24: 1 air-fuel ratio on part load; this reduces the combustion temperatures, and hence NO,, and also reduces the pumping work over a substantial part of the operating range. The power output of a diesel engine is controlled by changing the amount of fuel injected into the cylinder - the air quantity is not controlled. This means that the overall air-fuel ratio of the diesel engine changes with load, and the quality of the charge is controlled. This is referred to as qualitative governing and has the benefit of not requiring throttling of the intake charge. The diesel engine operates over a broad range of air-fuel ratios. However, the richest operating regime of a diesel engine is usually at an air- fuel ratio of more than 20 : 1, whereas the petrol engine can achieve the stoichiometric ratio of about 15 : 1. This means that, even at the same engine speed, the diesel engine will only produce about 70% of the power output of the petrol engine. However, the lack of throttling, higher compression ratio, and shorter combustion period mean that the diesel engine has a higher thermal efficiency than the petrol engine. (Note: the efficiency of engines should be compared on the basis of the specific fuel consumption [kg/kW h] rather than miles per gallon bscause the calorific values [kJ/kg] of diesel fuel and petrol are approximately the same, but the calorific value of diesel fuel in W/m3 is about 15% greater than that of petrol because the fuel density is about 15% greater. A diesel vehicle which achieves only 15% more miles/gallon than a petrol engine is not more efficient!) Gas turbine combustion chambers and boilers also control their output in a qualitative manner, by varying the air-fuel ratio.

An important feature of a flame is its rate of progress through the mixture, referred to as the flame speed. Flame speed is easiest understood when related to a premixed laminar flame, but the concept of flame speed applies in all premixed combustion.

15.2 Thermodynamics of combustion

When fuel and air are mixed together they are in metastable equilibrium, as depicted schematically in Fig 15.1. This means that the Gibbs energy of the reactants is not at the minimum, or equilibrium, value but is at a higher level. Fortunately it is still possible to apply the laws of thermodynamics to the metastable state, and hence the energies of reactants can be calculated in the usual way. Even though the reactants are not at the minimum Gibbs energy, spontaneous change does not usually occur: it is necessary to provide a certain amount of energy to the mixture to initiate the combustion - basically to ‘ionise’ the fuel constituents. Once the fuel constituents (usually carbon and hydrogen) have been ‘ionised’ they will combine with the oxygen in the air to form compounds with a lower

Page 309: Advanced Thermodynamics for Engineers

Thermodynamics of combustion 293

Gibbs energy than the reactants; these are called the products (usually carbon dioxide, water, etc). In the case of the spark-ignition engine the ionisation energy is provided by the spark, which ignites a small kernel of the charge from which the flame spreads. In the diesel engine part of the mixture produced in the cylinder cannot exist in the metastable state: it will spontaneously ignite. Such a mixture is termed a hypergolic mixture.

9 2 x e r3

Q) \ Metastable

\ Stable equilibrium

State Fig. 15.1 Energy states associated with combustion

Some mixtures are unstable at room temperature, and their constituents spontaneously ignite. An example of such a mixture is hydrogen and fluorine. This concept of spontaneous ignition will be returned to later.

All of these thermodynamic processes take place somewhere in the combustion zone, and many of them occur in the flame. Before passing on to the detailed discussion of flames it is worthwhile introducing some definitions and concepts.

15.2.1 REACTION ORDER

The overall order of a reaction is defined as n = cy=, v,, summed over the q species in the reactants.

First-order reactions

These are reactions in which there is spontaneous disintegration of the reactants. These reactions do not usually occur, except in the presence of an ‘inert’ molecule.

Second-order reactions

These are the most common reactions because they have the highest likelihood of a successful collision occurring.

Third-order reactions

These are less likely to occur then second-order ones but can be important in combustion. An example is when OH and H combine to produce an H,O molecule. This H,O molecule

Page 310: Advanced Thermodynamics for Engineers

294 Combustion and James

will tend to dissociate almost immediately unless it can pass on its excess energy - usually to a nitrogen molecule in the form of increased thermal energy.

15.2.2 PROCESSES OCCURRING IN COMBUSTION

The thermodynamics of combustion generally relates to the gas in isolation from its surroundings. However, the surroundings, and the interaction of the gas with the walls of a container, etc, can have a major effect on the combustion process. The mechanisms by which the combusting gas interacts with its container are:

(i) transport of gaseous reactant to the surface - diffusion; (ii) adsorption of gas molecules on surface; (iii) reaction of adsorbed molecules with the surface; (iv) desorption of gas molecules from the surface; (v) transport of products from the surface back into the gas stream - diffusion.

These effects might occur in a simple container, say an engine cylinder, or in a catalytic converter. In the first case the interaction might stop the reaction, while in the second one it might enhance the reactions.

' 15.3 Explosion limits

The kinetics of reactions was introduced previously, and it was stated that reactions would, in general, only occur when the atoms or ions of two constituents collided. The reaction rates were derived from this approach, and the Arrhenius equations were introduced. The tendency for a mixture spontaneously to explode is affected by the conditions in which it is stored. A mixture of hydrogen and oxygen at 1 bar and 500°C will remain in a metastable state, and will only explode if ignited. However, if the pressure of that mixture is reduced to around 10 mm Hg (about 0.01 bar) there will be a spontaneous explosion. Likewise, if the pressure increased to about 2 bar there would also be an explosion. It is interesting to examine the mechanisms which make the mixture become hypergolic. The variation of explosion limits with state for the hydrogen-oxygen mixture is shown in Fig 15.2.

lOUU0

h

r" E E , IO0 v

B 2 2 a

I U

I

3911 .rin 430 45(1 471 490 510 530 550 570 590

Temperature / ("C)

Fig. 15.2 Explosion limits for a mixture of hydrogen and oxygen (from Lewis and von Elbe, 1961)

Page 311: Advanced Thermodynamics for Engineers

Explosion limits 295

The kinetic processes involved in the H,-0, reaction are:

H+O,-OH

which leads to the further branching steps

0 + H 2 4 OH + H OH + H , 4 H 2 0 + H

(15.1)

The first two of the three steps are called branching steps and produce two radicals (highly reactive ions) for each one consumed. The third step does not increase the number of radicals. Since all steps are necessary for the reaction to occur, the multiplication factor (i.e. the number of radicals produced by the chain) is between 1 and 2. The first step is highly endothermic (it requires energy to be supplied to achieve the reaction), and will be slow at low temperatures. This means that an H atom can survive a lot of collisions without reacting, and can be destroyed at the wall of the container. Hence, H2-02 mixtures can exist in the metastable state at room temperatures, and explosions will only occur at high temperatures, where the first step proceeds more rapidly.

15.3.1 EXPLODE The effect of the multiplication factor can be examined in the following way. Assume a straight chain reaction has lo* collisions/s, and there is 1 chain particle/cm3, with 10'' molecules/cm3. Then all the molecules will be consumed in 10" seconds, which is approximately 30 years. However, if the multiplication factor is now 2 then 2" 10". giving N=62. This means that all the reactants' molecules will be consumed in 62 generations of collisions, giving a total reaction time of 62 x O-' seconds, or 0.62 ps: an extremely fast reaction! If the multiplication factor is only 1.01 then the total reaction time is still only 10 ms. Hence, the speed of the reaction is very dependent on the multiplication factor of the reactions, but the overall multiplication factors do not have to be very high to achieve rapid combustion.

THE EFFECT OF MULTIPLICATION FACTOR ON THE TENDENCY TO

A general branched chain reaction may be written

M - R initiation k l

k 2 R + M - a R + M chain branching, a > 1

k3 R + M - P product formation removes radical

chain termination I R 'L, destruction

R A destruction

wall

gas

(15.2)

where M is a molecule, R is a radical, and P is a product. a is the multiplication factor. The value of a necessary to achieve an explosion can be evaluated. The rate of formation of the product, P , is given by

-- d [ P 1 - k , [ R ] [ M ] dt

(15.3)

Page 312: Advanced Thermodynamics for Engineers

296 Combustion and flames

The steady state condition for the formation of radicals is

-- d[RI - 0 = k l [ M l + k 2 ( a - l ) [ R ] [ M ] - k 3 [ R ] [ M ] - k 4 [ R ] - k , [ R ] dt

Solving eqn (15.4) for R and substituting into eqn (15 .3) gives

(15 .4)

(15.5)

The rate of production of the product P becomes infinite when the denominator is zero, giving

(15.6)

Thus, if amact > aCnt the reaction is explosive; if the axact < a,fit then the combustion is non-explosive and progresses at a finite rate. A few explosion limits, together with flammability limits are listed in Table 15.1. This text will concentrate on non-explosive mixtures from now on.

Table 15.1 Flammability and explosion limits (mixtures defined in % volume) at ambient temperature and pressure

Lean Rich

Mixture Flammability Explosion Flammability Explosion Stoichiometric ~~~~~~

H, - air 4 18 74 59 29.8 co-0, 16 38 94 90 66.7 CO-air 12.5 74 29.8 NH3-02 15 25 79 75 36.4 c3Hfl-02 2 3 55 37 16.6 CH.-air 5.3 15 9.5 1 C,H, -air 3.0 12.5 5.66 C3H, -air 2.2 9.5 4.03 C,H,, - ax 1.9 8.5 3.13

15.4 Flames

A flame is the usual mechanism by which combustion of hydrocarbons takes place in air. It is the region where the initial breakdown of the fuel molecules occurs. There are two different types of flame, as described above: premixed flames and diffusion flames. Premixed flames will be dealt with first because it is easier to understand their mechanism.

15.4.1 PREMIXED FLAMES

Premixed flames occur in any homogeneous mixture where the fuel and the oxidant are mixed prior to the reaction. Examples are the Bunsen burner flame and the flame in most spark-ignited engines. Premixed flames can progress either as deflagration or detonation

Page 313: Advanced Thermodynamics for Engineers

Flames 297

processes. This text will consider only deflagration processes, in which the flame progresses subsonically. Detonation processes do occur in some premixed, spark-ignited engines, when the ‘end gas’ explodes spontaneously making both an audible knock and causing damage to the combustion chamber components.

When considering laminar flame speed, it is useful to start with a qualitative analysis of a Bunsen burner flame, such as depicted in Fig 15.3. If the flow velocity at the exit of the tube is low then the flow of mixture in the pipe will be laminar. The resulting flame speed will be the laminar flame speed. While most flames are not laminar, the laminar flame speed is a good indication of the velocity of combustion under other circums- tances. It can be seen from Fig 15.3(b) that the shape (or angle) of the inner luminous cone is defined by the ratio of the laminar flame speed (or burning velocity), ut, to the flow velocity of the mixture. In fact, the laminar flame speed ut= ug sin a. While this is a relatively simple procedure to perform, it is not a very accurate method of measuring laminar flame speed because of the difficulty of achieving a straight-sided cone, and also defining the edge of the luminous region. Other methods are used to measure the laminar flame speed, including the rate of propagation of a flame along a horizontal tube and flat burners.

. . :f

/

Premixed gas and air

(a) (b)

Fig. 15.3 Schematic diagram of Bunsen burner flame: (a) general arrangement; (b) velocity vectors

15.4.2 LAMINAR FLAME SPEED

There are a number of theories relating to laminar flame speed. These can be classified as:

0 thermal theories 0 diffusion theories 0 comprehensive theories.

The original theory for laminar flame speed was developed by Mallard and le Chatelier (1883), based on a thermal model. This has been replaced by the Zel’dovitch and Frank- Kamenetsky (1938), and Zel’dovitch and Semenov (1940) model which includes both

Page 314: Advanced Thermodynamics for Engineers

298 Combustion and flames

r -

I I I

____+ +--.' 3 .+-- +--a +-- +--%+-- Mass flow

I I I + - - lndicates thermal energy flow I , I I I

I I I

I I I

I I I Tb : Concentration I I I Temperature of reactants- I - - - - - -

I - \ I I I \

\ I I

I \ I I ',

Fig. 15.4 Schematic diagram of interactions in plane combustion flame

thermal and species diffusions across the flame. The details of these models will not be discussed but the results will simply be presented.

A plane flame in a tube may be shown schematically as in Fig 15.4. It can be seen that the thermal diffusion goes from right to left in this diagram, i.e. against the direction of flow. The flame also attempts to travel from right to left, but in this case the gas flow velocity is equal to the flame speed. If the gas had been stationary then the flame would have travelled to the left at the laminar flame speed.

Considering the physical phenomena occurring in the tube: heat flows, by conduction, from the burned products zone (b) towards the unburned reactants zone (u), while the gas flows from u to b. A mass element passing from left to right at first receives more heat by conduction from the downstream products than it loses by conduction to the reactants, and hence its temperature increases. At temperature T, the mass element now loses more heat to the upstream elements than it receives from the products, but its temperature continues to increase because of the exothermic reaction taking place within the element. At the end of the reaction, defined by Tbr the chemical reaction is complete and there is no further change in temperature.

The Zel'dovitch et al. (1938, 1940) analysis results in the following equation for laminar flame speed

(15.7)

In obtaining eqn (15.7) the assumption had been made that the Lewis number

L e = k / p c , D = l

where k = thermal conductivity, p = density, cp = specific heat at constant pressure, and D =mass diffusivity Hence, Le is the ratio between thermal and mass diffusivities, and

Page 315: Advanced Thermodynamics for Engineers

Flames 299

this obviously has a major effect on the transport of properties through the reaction zone. The assumption Le = 1 can be removed to give the following results for first- and second- order reactions.

For first-order reactions

and for second-order reactions

(15.8a)

(15.8b)

where 2' is the pre-exponential term in the Arrhenius equation and c, is the initial volumetric concentration of reactants.

Equations (15.8a) and (15.8b) can be simplified to

(15.9)

Hence the laminar flame speed is proportional to the square root of the product of thermal diffusivity, a , and the rate of reaction, R. Glassman (1986) shows that the flame speed can be written as

(1 5.10)

which is essentially the same as eqn (15.9), where R=Ze-EtRTb. Obviously the laminar flame speed is very dependent on the temperature of the products, Tb, which appears in the rate equation. This means that the laminar flame speed, ul, will be higher if the reactants temperature is high, because the products temperature will also be higher. It can also be shown that u1 = P ( ~ - ' ) / ' , where n is the order of the reaction, and n = 2 for a reaction of hydrocarbon with oxygen. This means that the effect of pressure on u1 is small. Figure 15.5 (from Lewis and von Elbe, 1961) shows the variation of u1 with reactant and mixture strength for a number of fundamental 'fuels'. It can be seen that, in general, the maximum value of uI occurs at close to the stoichiometric ratio, except for hydrogen and carbon monoxide which have slightly more complex reaction kinetics. It is also apparent that the laminar flame speed is a function both of the reactant and the mixture strength. The effect of the reactant comes through its molecular weight, m,. This appears in more than one term In eqn (15.10) because density and thermal conductivity are both functions of m,. The net effect is that ul= l/mw. This explains the ranking order of flame speeds shown in Fig 15.5, with the laminar flame speed for hydrogen being much higher than the others shown. While the molecular weight is a guide to the flame speed of a fuel, other more complex matters, such as the reaction rates, included as Z' in these equations, also have a big influence on the results obtained. Figure 15.6, from Metgalchi and Keck (1980, 1982) shows a similar curve, but for fuels which are more typical of those used in spark-ignition engines.

Page 316: Advanced Thermodynamics for Engineers

Fig. 155

Stoichiornetnc Stoichiornemc Stoichiometric ' H,,CO i

250 h m --.

E 200

3 r= 50 .-

5 4

0 0 10 20 30 40 50 60 70

% gas in air

Variation of laminar flame speed with reactants and mixture strength. p = 1 bar; Tu

50

45

2 25 f

E .-

15

1

=298 K

0.8 1 .o 1.2 1.4

Fuel-air equivale~ce ratio, B, Fig. 15.6 Variation of laminar flame speed with mixture strength for typical fuels (based on 1 a m ,

300 K)

Page 317: Advanced Thermodynamics for Engineers

Flames 301

It can be seen that the laminar flame speed is dependent on mixture strength, and this has a major influence on the design of engines operating with lean mixtures. The other feature to notice is that the laminar flame speed would remain approximately constant in an engine operating over a speed range of probably 800 to 6000 rev/min. If the combustion process depended on laminar burning then the combustion period in terms of crank angle would change by a factor of more than 7 : 1. Consider an engine operating with methane (CH,,) at stoichiometric conditions: the laminar flame speed is about 50 cm/s (0.5 m/s). If the engine bore is 100 mm then the combustion period will be 0.1 s. At 800 rev/min this is equivalent to 480" crankangle - longer than the compression and expansion periods! Obviously the data in Figs 15.5 and 15.6 are not directly applicable to an engine; this is for two reasons.

First, the initial conditions of the reactants, at temperatures of 298 K or 300 K, are much cooler than in an operating engine. Kuehl (1962) derived an expression for the combustion of propane in air, which gave

(15.11)

where

p = pressure (bar) T = temperature (K) u, = laminar flame speed (m/s)

It can be seen from eqn (15.1 1) that the effect of pressure on flame speed is very small, as suggested above. Figure 15.7 shows how the laminar flame speed increases with reactants temperature. It has been assumed that the adiabatic temperature rise remains constant at 2000 K, which is approximately correct for a stoichimetric mixture. It can be seen that the speed increases rapidly, and reaches a value of around 4 m/s when the reactants temperature

4.5 '1

300 400 500 600 700 800 900 1000

Reactants temperature / (K)

Fig. 15.7 Variation of laminar flame speed with reactants' temperature (predicted by Kuehl's equation with p = 1 bar)

Page 318: Advanced Thermodynamics for Engineers

302 Combustion andflames

is IO00 K. This is an increase of about a factor of 8 on the previous value which would reduce the combustion duration to about 60" crankangle. This is still quite a long combustion duration, especially since it has been evaluated at only 800 rev/min, some other feature must operate on the combustion process to speed it up. This, second, parameter is turbulence, which enhances the laminar flame speed as described below.

15.5.3 TURBULENT FLAME SPEED

It was shown previously that the laminar flame speed is too low to enable engines to operate efficiently, particularly if they are required to work over a broad speed range. The laminar flame speed must be enhanced in some way, and turbulence in the flow can do this. A popular model describing how turbulence increases the flame speed is the wrinkled laminarflame model, which is shown diagrammatically in Fig 15.8.

Burned gas

Unburned

Lamina flame

Turbulent (wrinkled) flame

Fig. 15.8 Comparison of laminar and turbulent flames

The effect of turbulence on the flame is threefold:

0 the turbulent flow distorts the flame so that the surface area is increased; 0 the turbulence may increase the transport of heat and active species; 0 the turbulence may mix the burned and unburned gases more rapidly.

The theory of turbulent flames was initiated by Damkohler (1940) who showed that the ratio of turbulent to laminar flame speeds, based on large scale eddies in the flow, is

( 1 5.1 2)

where

E = eddy diffusivity Y = kinematic viscosity of the unburned gas

Heikal et al. (1979) applied a similar approach to engine calculations, and defined the ratio of turbulent to laminar flame speed, often called the flame speed factor, as

(1 5.13)

Page 319: Advanced Thermodynamics for Engineers

Flammability limits 303

where a = molecular thermal diffusivity P, = Prandtl number vp = mean piston speed a , b , c , dare all empirical constants

Experience shows that in spark ignition engines the flame speed factor is a strong function of engine speed but is not greatly affected by load. Experiments by Lancaster et al. (1976) have also shown that the level of turbulence intensity in an engine cylinder increases with engine speed, but is not quite proportional to it. This means that while turbulence increases the flame speed significantly, the length of the burning period increases as engine speed is increased, which explains why the ignition timing has to be advanced.

15.5 Flammability limits

Flammability limits were introduced in Table 15.1, where they were listed with the explosion limits. The flammability limit of a mixture is defined as the mixture strength beyond which, lean or rich, it is not possible to sustain a flame. The flammability limit, in practice, is related to the situation in which the flame is found. If the flame is moving in a confined space it will be extinguished more easily because of the increase in the interaction of the molecules with the walls; this is known as quenching. It is also possible for the flame to be extinguished if the level of turbulence is too high, when the flame is stretched until it breaks. The lean flammability limit is approximately 50% of stoichiome- tric, while the rich limit is around three times stoichiometric fuel-air ratio.

Bradley et al. (1987, 1992) investigated the way in which turbulence ‘stretches’ the flame and causes it to be extinguished. The results are summarised in Fig 15.9, which is a graph of the ratio of turbulent to laminar flame speed, i.e. the flame speed factor defined in eqn (15.13), against the ratio of turbulence intensity to laminar flame speed. The abscissa is closely related to the Karlovitz number, K , which is a measure of the flame stretch:

(1 5.14)

where F, = area of laminar flame; z, = transit time for flow through laminar flame; u‘ = turbulence intensity; 1, = Taylor microscale; 6, = laminar flame thickness; u, = laminar flame speed.

Also shown in Fig 15.9 are lines of constant KLe, which is the product of Karlovitz and Lewis numbers, and lines of constant Re/Le2, which is the ratio of Reynolds number to the square of Lewis number. If the value of KLe is low (e.g. 0.005) then the flame is a wrinkled laminar one, while if the value of KLe is high ( e g 6 ) the flame is stretched

Page 320: Advanced Thermodynamics for Engineers

304 Combustion and flames

sufficiently to quench it. This shows that the design of a combustion chamber must be a compromise between a high enough value turbulence intensity to get a satisfactory flame speed, and one which does not cause extinction of the flame.

2 0 -1

0

5 - c

.- 0 I

0

e:

I 8

16

14

12

10

8

6

4

2

0 63

Disrupted, partial quench region I

Wrinkled 6 laminar region

Quench region

- Constant KLe

- r o n s t a n t Re/Le2

0 I

0 2 4 6 8 10 12 14 16 18 20

Ratio ( turbulence intens i ty / laminar f l a m e speed)

I ' l ' l ' l ~ l ' l ' l ' l ' l ' l

Fig. 15.9 The effect of turbulence on the turbulent flame speed and tendency to quench for a premixed charge

It can be seen from Table 15.1 that the spread of flammability for hydrogen is much higher than the other fuels This makes it an attractive fuel for homogeneous charge engines because it might be possible to control the load over a wide range of operation by qualitative governing rather than throttling. This would enable the hydrogen powered engine to achieve brake thermal efficiencies similar to those of the diesel engine. The restricted range of flammability limits for hydrocarbon fuels limits the amount of power reduction that can be achieved by lean-burn running; it also restricts the ability of operating the engine lean to control NO,. Honda quote one of their engines operating as lean as 24: 1 air-fuel ratio, which enables both the engine power to be reduced and the emissions of NO, to be controlled. Such lean-bum operation is achieved through careful design of the intake system and the combustion chamber. In practice, in a car engine the lean limit is set by the driveability of the vehicle and the tendency to misfire. A small percentage of misfires from the engine will make the uHC emissions unacceptable - these misfires might not be perceptible to the average driver.

The rich and lean flammability limits come closer together as the quantity of inert gas added to a mixture is increased. Fortunately, the rich limit of flammability is more affected than the lean one, and basically the lean limit, which is usually the important one for engine operation, is not much changed.

15.6 Ignition

Tne ignition process is an extremely important one in the homogeneous charge engine because it has to be initiated by an external source of energy - usually a spark-plug. It can be shown that the minimum energy for ignition, based on supplying sufficient energy to the volume of mixture in the vicinity of the spark-gap to cause a stable flame, is

Page 321: Advanced Thermodynamics for Engineers

Difisionflames 305

Q p , m i n ~ k3/p2u:. This means that the minimum energy is increased as the inverse of the laminar flame speed cubed. In other words, much more energy is required when u, is low. The typical level of energy required to ignite the mixture in a spark-ignition engine is around 30 ml. As shown above, ut is dependent on the initial temperature of the reactants, and the equivalence ratio of the charge (because this affects the adiabatic temperature rise, and hence TJ. This means that the strongest spark is needed when the engine is cold, and also if the engine is running lean. This dependence also explains why it is attractive to develop stratified charge engines, with a rich mixture zone around the spark-plug. The typical variation of minimum energy to ignite a mixture is shown in Fig 15.10.

Air-fuel ratio

Fig. 15.10 Minimum ignition energy against air-fuel ratio

15.7 Diffusion flames

The previous sections all related to premixed flames of the type found in Bunsen burners, or spark-ignition engines. The other major class of flames is called diffusion flames; in these flames the rates of reaction are not controlled by the laminar flame speed but the rate at which the fuel and air can be brought together to form a combustible mixture. This type of combustion occurs in

0 open flames, when mixing with secondary air enables combustion of a rich premixed core to continue to completion;

0 gas turbine combustion chambers, when the liquid fuel sprays are mixed with the air in the combustion chamber;

0 diesel engines, when the injected fuel has to mix with the air in the chamber before combustion can take place.

A typical arrangement of a diffusion flame might be that shown in Fig 15.11. This is a simple example of a jet of hydrogen passing into an oxygen atmosphere. The principle is the same if the fuel is a more complex gaseous one passing into air; the major difference will be that the products of combustion will be more complicated. The three sections across the jet (at A, B and C) show the way in which the oxygen diffuses into the jet, usually by turbulent mixing brought about by the jet entraining the surrounding oxygen. At section A the hydrogen and oxygen are completely separate, as indicated by the mass fraction curves. By section B, some way from the end of the nozzle, the hydrogen has been mixed with the oxygen just outside the jet diameter due to turbulent entrainment.

Page 322: Advanced Thermodynamics for Engineers

306 Combustion andflames

There has not yet been any mixing of hydrogen and oxygen in the potential core. Some combustion has also taken place by this section, as indicated by the mass fraction of water. Section C is located almost at the end of the mixing zone, after the end of the potential core, and there is no pure hydrogen left in the jet. The edge of the mixing zone is now well within the diameter of the jet.

Considering the concentration of hydrogen and oxygen along the centre-line of the jet, it can be seen that there is pure hydrogen right up to the end of the potential core. After that the hydrogen and oxygen on the centre-line combine to form water, and it is not until the end of the mixing zone that the oxygen concenwation starts to rise again, as the water and oxygen mix and dilute each other.

In this example it has been assumed that the oxygen and hydrogen burn as soon as they come intimately in contact. This presupposes that the chemical reaction rate is much faster than the diffusion rates; this is usually a reasonable assumption.

A similar, but more complex analysis may be made of the injection of diesel fuel into the cylinder of a diesel engine. In this case the entrainment of the fuel and air does not take place in the gaseous phase, but occurs because the droplets of fuel leaving the nozzle impart their momentum on the surrounding air by aerodynamic drag. Once the fuel and air are mixed it is possible for the droplets to evaporate to create a combustible mixture. The details of diesel combustion are beyond this course, but the principles are similar to the combustion of gaseous jets. Gas turbine combustion chambers work in a similar way, with a spray of fuel entraining primary air to initiate the combustion, and subsequent entrainment of secondary air to complete the process. Individual combustion systems will now be considered.

Fig. 15.11 Combustion of a jet of hydrogen in an oxygen atmosphere

Page 323: Advanced Thermodynamics for Engineers

Engine combustion systems 307

15.8 Engine combustion systems Different engines have different combustion systems which reflect the way in which the fuel and air are brought together. These systems will be described briefly.

15.8.1 SPARK-IGNITION ENGINES

In spark-ignition engines the fuel and air are usually premixed prior to admission to the engine cylinder. This is done either in a carburettor, or more commonly now by a port or manifold fuel injection system. [Some of the two- and four-stroke engines being investigated for use in cars do have direct injection into the engine cylinder using an air- blast injector. These engines will not be considered here, but are a hybrid between the fully premixed spark-ignition engine and the diesel engine in their combustion mechan- ism.] All these systems prepare the charge prior to it entering the cylinder, although it is probable that the fuel enters the cylinder with a large proportion in the liquid phase. Under fully warmed-up conditions this fuel will have evaporated by the time of ignition. At start- up this will not be the case, and enrichment, beyond stoichiometric, is done to ensure that the light fractions of the fuel give a combustible mixture; the remaining liquid fuel causes high levels of unburned hydrocarbons (uHCs). It was stated above that a high level of turbulent gas motion in the cylinder will increase the flame speed, and this can be achieved by various mechanisms. In older engines the shape of the piston and cylinder head produced a squish motion as the piston approached tdc, which enhanced the turbulence in the region of the spark plug, and increased the flame speed (see Fig 15.12).

(a) (b)

Fig. 15.12 Squish flow in ‘bath tub’ combustion chamber: (a) piston at mid-stroke; (b) piston approaching tdc and squishing gas from top land region

Other designs have been proposed, including the May combustion chamber (Figs 15.13(a) and (b)) which produces a high level of turbulence by ‘squeezing’ the gas into a small combustion chamber under either the intake or exhaust valve. While this system produces high turbulence it occurs too late to achieve its aims. A May chamber was fitted to a Jaguar engine and produced better fuel economy by enabling leaner mixtures to be used. More modem engines attempt to increase the turbulence levels around the spark plug by the break-up of barrel swirl or tumble. The gas entering the engine has a combination of swirl (vortex motion in a horizontal plane) and barrel swirl (vortex motion

Page 324: Advanced Thermodynamics for Engineers

308 Combustion andflames

in the vertical plane) (see Figs 15.13 (c) and (d)). Swirl momentum is preserved during compression, but is not very useful for spark-ignition engine combustion. Barrel swirl cannot be preserved as ordered motion because the shape of the vortex is destroyed as the aspect ratio of the combustion chamber changes: barrel swirl is broken down into smaller scale motion which enhances the flame speed.

Fig. 15.13 High activity cylinder heads for spark-ignition engines: (a) May combustion chamber - note combustion chamber under valve; (b) May combustion chamber with piston near tdc - note squishing of charge into compact chamber; (c) modem pent-roof (4 valve) head - note barrel swirl set up in cylinder; (d) pent-roof chamber with piston near tdc - note high activity of gas around

Plug

Further features which must be borne in mind with spark-ignition engine combustion cllamkrs are:

(i) reducing zones where combustion can be quenched (see top land region in Fig 15.11);

Page 325: Advanced Thermodynamics for Engineers

Engine combustion systems 309

(ii) limiting zones which might trap the end gases and cause detonation; (iii) reducing crevice volumes where uHCs might be trapped, e.g. around top ring

groove.

15.8.2 DIESEL ENGINES

The design of diesel engine combustion chambers is different from that of spark-ignition engines because of the nature of the diesel process. Fuel is injected into the diesel engine cylinder, in liquid form, through a high pressure injector. This fuel enters the engine as a jet, or jets, which has to entrain air to enable evaporation of the fuel and subsequent mixing to a point where hypergolic combustion occurs. The mixing and combustion processes are similar to those shown in Fig 15.11 for a gaseous jet. The droplet size of the fuel varies but is of the order of 20 pm; the size depends on the injection nozzle hole size and the fuel injection pressure, which might be 0.20 mm and 700 bar in a small high-speed direct injection diesel engine. The prime considerations in the design of a combustion chamber for a diesel engine are to obtain efficient mixing and preparation of the fuel and air in the time available in the cycle.

The majority of small diesel engines for passenger cars use the indirect injection (idi) process. In these engines, see Fig 15.14(b), the fuel is injected not directly into the main combustion chamber but into a small swirl chamber (e.g. Ricardo Comet V) connected to the main chamber by a throat. This approach was adopted because it was possible to achieve good mixing of the fuel and air in the pre-chamber due to the high swirl velocities that could be generated there as air was forced into the pre-chamber by the piston travelling towards tdc. This approach had a further advantage that relatively simple fuel injection equipment could be used, with a single hole nozi..le and low preysure ;njection pump. Combustion of the rich mixture commences in the pre-chamber, and the burning gases enter the main chamber (which contains pure air) and generate high turbulence which ensures good mixing of the burning plume and the ‘secondary’ air. Such combustion systems are very tolerant of fuel quality, need relatively simple fuel injection equipment,

Fuel injector

Fuel injector Swirl chamber

(a) (b)

Fig. 15.14 Some combustion chamber arrangements for diesel engines: (a) basic schematic of direct injection diesel engine; (b) basic schematic of indirect injection diesel engine

Page 326: Advanced Thermodynamics for Engineers

3 10 Combustion andflames

and can be run at relatively low air-fuel ratios before producing black smoke. Their disadvantage is that their fuel economy is about 10% or more worse than their direct injection counterparts.

Direct injection (di) diesel engines (shown schematically in Fig 15.14(a)) dominate the larger size range. Nearly all truck diesel engines are direct injection, as are those for rail and marine applications. The bottom limit of size for di diesel engines is reducing all the time, and four cylinder engines as small as 2 litres are available for van and car applications. The largest di diesel engines have quiescent combustion chambers, in which there is no organised air motion. The mixing of the fuel and air is achieved by the multiple fuel jets entraining air into themselves and bringing about the necessary mixing. More than six holes might be used in the fuel injector nozzle to give good utilisation of the air in the chamber. The relatively low engine speed of large engines allows sufficient time for combustion to occur. As the engine bore size reduces so the engine speed increases, and the time available for combustion becomes shorter. Also, the appropriate size of injector hole reduces, until it reaches the limit that can be achieved by production techniques (around 0.18 mm diameter). The increased engine speed, and reduced time for mixing, requires that the rate of mixing of the fuel and air is enhanced above that which can be achieved using a quiescent combustion chamber. The mixing rate can be increased by imparting air motion to the charge, and this is done in the form of swirl in the diesel engine. The fuel jets injected into a swirling flow have the peripheral fuel stripped off to form a combustible mixture, which is where ignition is initiated. Figure 15.15 shows the ignition points located by high speed photography in a high speed di engine.

Fig. 15.15 Sprays and combustion initiation points in a small di diesel engine: H = high swirl; M = medium swirl; L = low swirl

Figure 15.15 also shows that the fuel jets impact upon the wall of the piston bowl. This also plays an important contribution in the mixing of the fuel and air, and is necessary at present because high enough mixing rates cannot be achieved in the centre of the bowl itself.

Page 327: Advanced Thermodynamics for Engineers

Engine combustion systems 31 1

The design of combustion systems for diesel engines must aim at ‘optimising’ the overall combustion process. This takes place in two distinct modes: pre-mixed and diffusion combustion. The first fuel that is injected into the cylinder is not ready to ignite until it has evaporated and produced the necessary conditions for hypergolic combustion. This takes a finite time, referred to as the ignition de2ay period, during which the fuel is prepared but not yet ignited. At the end of this period there is rapid combustion of the premixed fuel and air, which gives rise to a high rate of heat release and produces high temperatures in the combustion chamber. This period has a major effect on the amount of NO, produced in the engine. Typical equations for the ignition delay period are

(15.15a)

or [EA( l /RT- 1/17190)(21.2/(p - 12.4))0.631

ti, = (0.36 + 0.22Vp)e deg crankangle ( 1 5.15b)

where Vp = mean piston speed (m/s) p = pressure (bar) T =temperature (K)

E , = activation energy = 618840/(CN + 25) CN = cetane number

Both eqns (15.15a) and (15.15b) have a similar form, and are related to the Arrhenius equation introduced in Chapter 14 to define the rate equations. Equation (15.15b) is a more recent formulation than eqn (15.15a), and has a more complex structure. Both equations are the result of experimental tests on engines with a range of fuels, and cannot be extended far beyond the regime under which they were evaluated, but they do give a basic structure for ignition delay. It should be noted that eqn (15.15b) contains a term for the cetane number of the fuel. The value of E, reduces as the cetane number increases and this means that the ignition delay is inversely related to cetane number. A mechanical method for limiting the overall ignition delay is to use two-stage or split injection. In this type of system a small quantity of pilot fuel is injected into the cylinder some time prior to the main injection process. The pilot charge is prepared and ready to ignite before the main charge enters the chamber, and in this way the premixed combustion is limited to the pilot charge.

After the premixed period is over, the main combustion period commences, and this is dominated by diffusion burning, controlled by the mixing of the fuel and air. Whitehouse and Way (1970) attempted to model both periods by the two equations given below.

Reaction rate:

( 15.16a)

and preparation rate:

P = K’mi(’-x)mi( Po*)’ ( 15.16b)

Page 328: Advanced Thermodynamics for Engineers

3 12 Combustion and flames

where

pq = partial pressure of oxygen P = rate of preparation of fuel by mixing R = rate of reaction

mi = mass of fuel injected mu = mass of fuel unburned

These equations result in instantaneous heat re-sase patterns o the form shown in Fig 15.16. One of the diagrams has a short ignition delay, and it can be seen that the instantaneous rate of heat release does not reach such a high level as for the long delay. This is because the time for the physical and chemical processes to enable the fuel to reach a hypergolic state is less, and consequently less fuel is available for spontaneous ignition. The long delay results in a large amount of fuel burning spontaneously, with high temperature rises, high rates of pressure rise (dplda) and a high level of noise generation. This initial period is governed by the rate of reaction, R. After the premixed phase has taken place, the temperatures inside the combustion chamber are high and the rate of reaction is much faster than the rate of preparation, P. At this stage the combustion process is governed by eqn (15.16b), which models a diffusion process. During this process the rate of combustion is controlled by the rate at which the fuel and air mix, and in this phase the hydrocarbon fuel in the centre of the jet is burning in insufficient oxygen. The fuel pyrolyses and forms the precursors of the carbon particles produced in the exhaust system. It is important to mix the burning fuel with the air at the appropriate rate to ensure that the carbon produced during the combustion process is consumed before the exhaust valve opens.

a , Crankangle, a

Fig. 15.16 Effect of ignition delay on rate of heat release diagram

15.8.3 GAS TURBINES

Combustion in gas turbines takes place at essentially constant pressure, although there is a small pressure drop through the chamber which should be taken into account when

Page 329: Advanced Thermodynamics for Engineers

Engine combustion systems 313

undertaking design. The overall air-fuel ratio in the combustion chamber will be around 40: 1, although there will be regions in which much richer mixtures exist. Figure 15.17 is a schematic section of an annular combustion chamber, which will be located circumferentially around the body of the gas turbine. Annular combustion chambers have advantages over the older types of tubular combustors in terms of pressure loss and compactness.

I Dilubon Nozzle

Snout

I 1 Intermediate

hole Diluhon hole

Fig. 15.17 Section through aero gas turbine annular combustion chamber

The combustion in the chamber takes dace in a number of regions. When considering the -,ombustion pmcess it is useful to remember that the three parameters - temperature, turbulence, time - have to be optimised to achieve good, clean combustion. The air enters the chamber through the diffuser, where its velocity is reduced and the air-flow is distributed between the primary air entering tht cnout and the secondary air entering the inner and outer annuli. The snout imparts a high swirl into the flow to induce a strong recirculation region in the primary zone; this swirl enables the residence time in the primary zone to be lengthened without making this region excessively long. It also promotes good mixing through turbulence. The fuel injector will be either a high pressure spray atomiser or an air-blast injector. Sometimes a smaller pilot injector is used to assist ignition. Ignition is achieved by means of a powerful plasma jet igniter. In the primary combustion zone the air-fuel ratio will be approximately stoichiometric (15 : l), and the adiabatic temperature rise will take the temperature up to about 2000 K, at which temperature NO, will be formed. The gas will leave the primary zone and then be diluted with secondary air bleeding through the side walls of the chamber: the charge is then in the intermediate zone. This air will induce further turbulence and supply more oxygen to complete the burning of the soot formed during the rich primary combustion. The secondary air will also lower the gas temperature to about 1800 K, and reduce the amount of NO, produced. Next, the gases will enter the dilution zone where they will be further diluted, and their temperature will be reduced to about 1300-1400 K. The design of the dilution zone will also reduce the circumferential temperature distribution. Finally the cooling air which has been passed down the outside of the combustion chamber will be mixed in with the hot gases, and the products will be passed to the turbine through the nozzle, whch matches the flow from the combustion chamber to the requirements of the turbine.

Page 330: Advanced Thermodynamics for Engineers

314 Combustion and flames

15.9 Concluding remarks

The physical phenomena which affect combustion have been introduced. The differences between homogeneous, pre-mixed reactants and heterogeneous reactants, and their effect on combustion, have been described. It has been shown that pre-mixed combustible mixtures exhibit a characteristic combustion velocity called the laminar flame speed. The laminar flame speeed is related to the mixture strength and the temperature of the reactants, but is too slow for most engineering applications and must be enhanced by turbulence to achieve levels appropriate for powerplant. The level of turbulence must be balanced against the other combustion properties to ensure that the flame is not extinguished.

Diffusion flames which occur in heterogeneous mixtures have been described, and it has been shown that these rely on the mixing of the fuel and oxidant to achieve combustion of the mixture.

Examples have been given of combustion chambers for petrol and diesel engines, and gas turbines. The general principles derived in this chapter have been used to analyse their operation.

PROBLEMS 1 (a) One of the main problems encountered in the design of a diesel engine combustion

system is the mixing of the air and fuel sufficiently rapidly to ensure complete combustion. Explain, using diagrams, how these problems are catered for in the design of

(i) large automotive diesel engines; (ii) the smallest automotive diesel engines.

Compare and contrast the combustion systems of diesel and spark ignition Give two relative advantages of each type of combustion system. (b) engines in the forms they are applied to passenger cars.

2 (a) What is meant by the terms

(i) a global reaction; (ii) an elementary reaction; (iii) a reaction mechanism.

(b) Describe the steps required to form a chain reaction and explain why chain reactions are important in combustion. (c) A reaction is found to be 25 times faster at 400 K than at 300 K. Measurements of the temperature exponent yielded a value of 0.7. Calculate the activation energy of the reaction. How much faster will the reaction be at 1000 K?

3 A combustible mixture of gas and air is contained in a well insulated combustion bomb. It is ignited at a point and a thin flame propagates through the mixture, completely burning the reactants. This mechanism produces multiple zones of products. Prove that the temperature of an element of gas mixture which burned at pressure pb has a temperature T ( p, Pb) at a pressure p > Pb:

Page 331: Advanced Thermodynamics for Engineers

Problems 315

where

K = ratio of specific heats, QL = calorific value of fuel, and suffix 1 defines the conditions before ignition.

Calculate the final pressure, p z , in terms of pl, T I , QL, c,, and K. What is the difference between the final temperature and that of the first gas to bum if Tl = 300 K, K = 1.3 and QL/c, = 1500 K.

4 The structure of ethylene is H2C = CH,. Estimate the enthalpy of reaction when 1 kmol of ethylene is completely oxidised. Compare the value obtained with the tabulated value of - 1323.2 MJ/kmol. Give reasons for the difference between the values.

Neglecting dissociation, find the temperature reached after constant pressure combustion of ethylene with 50% excess air if the initial temperature of the reactants is 400 K. The specific heat at constant pressure, cp, of ethylene is approximately 1.71 kJ/ kg K over the temperature range of the reactants.

[1302.3 MJ/kmol; 2028 K]

5 A method of improving engine fuel consumption and reducing the emissions of NO, in a spark ignition engine is to run it lean, i.e. with a weak mixture. Discuss the problems encountered when running engines with weak mixtures, and explain how these can be overcome by design of the engine combustion chamber.

6 Describe the construction of a boiler for burning pulverised coal. Explain how this design optimises the temperature, turbulence and time required for good combustion. What are the main emissions from this type of plant, and how can they be reduced?

Page 332: Advanced Thermodynamics for Engineers

16 I r reve rsi bl e The rmod yna m ics

16.1 Introduction

Classical thermodynamics deals with transitions from one equilibrium state to another and since it does not analyse the changes between state points it could be called thermostatics. The term thermodynamics will be reserved, in this chapter, for dynamic non-equilibrium processes.

In previous work, phenomenological laws have been given which describe irreversible processes in the form of proportionalities, e.g. Fourier’s law of heat conduction, Ohm’s law relating electrical current and potential gradient, Fick’s law relating flow of matter and concentration gradient etc. When two of these phenomena occur simultaneously they interfere, or couple, and give rise to new effects. One such cross-coupling is the reciprocal effect of thermoelectricity and electrical conduction: the Peltier effect (evolution or absorption of heat at a junction due to the flow of electrical current) and thermoelectric force (due to maintenance of the junctions at different temperatures). It is necessary to formulate coupled equations to deal with these phenomena, which are ‘phenomenological’ inasmuch as they are experimentally verified laws but are not a part of the comprehensive theory of irreversible processes.

It is possible to examine irreversible phenomena by statistical mechanics and the kinetic theory but these methods are on a molecular scale and do not give a good macroscopic theory of the processes. Another method of considering non-equilibrium processes is based on ‘pseudo-thermostatic theories’. Here, the laws of thermostatics are applied to a part of the irreversible process that is considered to be reversible and the rest of the process is considered as irreversible and not taken into account. Thomson applied the second law of thermostatics to thermoelectricity by considering the Thomson and Peltier effects to be reversible and the conduction effects to be irreversible. The method was successful as the predictions were confirmed by experiment but it has not been possible to justify Thomson’s hypothesis from general considerations.

Systematic macroscopic and general thermodynamics of irreversible processes can be obtained from a theorem published by Onsager (1931a,b). This was developed from statistical mechanics and the derivation will not be shown but the results will be used. The theory, based on Onsager’s theorem, also shows why the incorrect thermostatic methods give correct results in a number of cases.

Page 333: Advanced Thermodynamics for Engineers

Entropy flow and entropy production 3 17

16.2 Definition of irreversible or steady state thermodynamics

All previous work on macroscopic ‘thermodynamics’ has been related to equilibrium. A system was said to be in equilibrium when no spontaneous process took place and all the thermodynamic properties remained unchanged. The macroscopic properties of the system were spatially and temporally invariant.

System 2

: System boundary

Distance

Fig. 16.1 Steady state conduction of heat along a bar

Consider the system shown in Fig 16.1, in which a thermally insulated rod connects two reservoirs at temperatures T2 and T, respectively. Heat flows between the two reservoirs by conduction along the rod. If the reservoirs are very large the conduction of energy out of, or into, them can be considered not to affect them. A state will be achieved when the rate of heat flow, dQ/dt, entering the rod equals the rate of heat flow, -dQ/dt, leaving the rod. If a thermometer were inserted at any point in the rod the reading would not change with time, but it would be dependent on position. If the rod were of uniform cross-section the temperature gradient would be linear. (This is the basis of the Searle’s bar conduction experiment. )

The temperature in the bar is therefore a function of position but is independent of time. The overall system is in a ‘stationary’ or ‘steady’ state but not in ‘equilibrium’, for that requires that the temperature be uniform throughout the system. If the metal bar were isolated from all the influences of the surroundings and from the heat sources, i.e. if it is made an isolated system, the difference between the steady state and equilibrium becomes obvious. In the case where the system was in steady state, processes would occur after the isolation (equalisation of temperature throughout the bar); where the system was already in equilibrium, they would not.

16.3 Entropy flow and entropy production

Still considering the conduction example given above. If the heat flows into the left-hand end of the bar due to an infinitesimal temperature difference, i.e. the process is reversible, the left-hand reservoir loses entropy at the rate

(16.1)

Page 334: Advanced Thermodynamics for Engineers

3 18 Irreversible thermodynamics

Similarly the right-hand reservoir gains entropy at the rate

1 dQ - - - -- - dt i’; dt

Thus, the total change of entropy for the whole system is

(16.2)

(16.3)

Now T2 > T, and therefore the rate of change of entropy, dS/dt > 0. To understand the meaning of this result it is necessary to consider a point in the bar.

At the point I from the left-hand end the thermometer reading is T. This reading is independent of time and is the reading obtained on the thermometer in equilibrium with the particular volume of the rod in contact with it. Hence, the thermometer indicates the ‘temperature’ of that volume of the rod. Since the temperature is constant the system is in a ‘steady state’, and at each point in the rod the entropy is invariant with time. However, there is a net transfer of entropy from the left-hand reservoir to the right-hand reservoir, i.e. entropy is ‘flowing’ along the rod. The total entropy of the composite system is increasing with time and this phenomenon is known as ‘entropy production’.

16.4 Thermodynamic forces and thermodynamic velocities

It has been suggested that for systems not far removed from equilibrium the development of the relations used in the thermodynamics of the steady state should proceed along analogous lines to the study of the dynamics of particles, i.e. the laws should be of the form

J = L X (16.4)

X is the thermodynamic force; L is a coefficient independent of X and J and is scalar in form, while both J and X are vector quantities.

where J is the thermodynamic velocity or flow;

The following simple relationships illustrate how this law may be applied. Fourier’s equation for one-dimensional conduction of heat along a bar is

(16.5)

where Q = quantity of energy (heat); T = temperature; A = area of cross-section; 1 = length; k = thermal conductivity.

Ohm’s Law for flow of electricity along a wire, which is also one-dimensional, is

(16.6)

where I = current; qr = charge (coulomb); e = potential difference (voltage); A = area of cross-section of wire; I = length; A = electrical conductivity.

Page 335: Advanced Thermodynamics for Engineers

Onsager's reciprocal relation 319

Fick's Law for the diffusion due to a concentration gradient is, in one dimension

dCi dn, dt dl

- = - k - (16.7)

where ni = amount of substance i , Ci = concentration of component i , and k = diffusion coefficient. This law was derived in relation to biophysics by analogy with the laws of thermal conduction to explain the flow of matter in living organisms. It will be shown later that this equation is not as accurate as one proposed by Hartley, in which the gradient of the ratio of chemical potential to temperature is used as the driving potential.

Other similar relationships occur in physics and chemistry but will not be given here. The three equations given above relate the flow of one quantity to a difference in potential: hence, there is a flow term and a force term as suggested by eqn (16.4). It will be shown that although eqns (16.5), (16.6) and (16.7) appear to have the correct form, they are not the most appropriate relationships for some problems. Equations (16.5), (16.6) and (16.7) also define the relationship between individual fluxes and potentials, whereas in many situations the effects can be coupled.

16.5 Onsager's reciprocal relation

If two transport processes are such that one has an effect on the other, e.g. heat conduction and electricity in thermoelectricity, heat conduction and diffusion of gases, etc., then the two processes are said to be coupled. The equations of coupled processes may be written

Equation (16.8) may also be written in matrix form as

(16.8)

(16.8a)

It is obvious that in this equation the basic processes are defined by the diagonal coefficients in the matrix, while the other processes are defined by the off-diagonal terms.

Consider the situation where diffusion of matter is occurring with a simultaneous conduction of heat. Both of these processes are capable of transferring energy through a system. The diffusion process achieves this by mass transfer, i.e. each molecule of matter carries some energy with it. The thermal conductivity process achieves the transfer of heat by the molecular vibration of the matter transmitting energy through the system. Both achieve a similar result of redistributing energy but by different methods. The diffusion process also has the effect of redistributing the matter throughout the system, in an attempt to achieve the equilibrium state in which the matter is evenly distributed with the minimum of order (i.e. the maximum entropy or minimum chemical potential). It can be shown, by a more complex argument, that thermal conduction will also have an effect on diffusion. First, if each process is considered in isolation the equations can be written

J , = L , , X ,

J 2 = L2,X2

the equation of conduction without any effect due to diffusion

the equation of diffusion without any effect due to conduction

Page 336: Advanced Thermodynamics for Engineers

320 Irreversible thermodynamics

Now, the diffusion of matter has an effect on the flow of energy because the individual, diffusing, molecules carry energy with them, and hence an effect for diffusion must be included in the term for thermal flux, J , . Hence

J1= LIlX, + LI2X2 (16.9)

where L I Z is the coupling coeficient showing the effect of mass transfer (diffusion) on energy transfer.

In a similar manner, because conduction has an effect on diffusion, the equation for mass transfer can be written

(16.10) 52 = L21 Xl + L22X2 where L,, is the coupling coefficient for these phenomena.

A general set of coupled linear equations is

(16.11) k

The equations are of little use unless more is known about the forces X, and the coefficients Lip This information can be obtained from Onsager’s reciprocal relation. There is considerable latitude in the choice of the forces X, but Onsager’s relation chooses the forces in such a way that when each flow Ji is multiplied by the appropriate force Xi, the sum of these products is equal to the rate of creation of entropy per unit volume of the system, 8, multiplied by the temperature, T. Thus

TO = JIXl + J2X2 + ... = C A X i (16.12) i

Equation (16.12) may be rewritten

(16.13)

where

Onsager further showed that if the above mentioned condition was obeyed, then, in

( 1 6.14)

This means that the coupling marrix in eqn (16.8a) is symmetric, Le. L,, = L,, for the particular case given above. The significance of this is that the effects of parameters on each other are equivalent irrespective of which is judged to be the most, or least, significant parameter. Consideration will show that if this were not true then it would be possible to construct a system which disobeyed the laws of thermodynamics. It is not proposed to derive Onsager’s relation, which is obtained from molecular considerations - it will be assumed to be true.

In summary, the thermodynamic theory of an irreversible process consists of first finding the conjugate fluxes and forces, Ji and xi , from eqn (16.13) by calculating the entropy production. Then a study is made of the phenomenological equations (16.11) and

general

Lik = Lki

Page 337: Advanced Thermodynamics for Engineers

The calculation of entropy production or entropyflow 321

Onsager’s reciprocal relation (16.14) is used to solve these. The whole procedure can be performed within the realm of macroscopic theory and is valid for any process.

16.6 The calculation of entropy production or entropy flow

In section 16.3 the concept of entropy flow was introduced. At any point in the bar, I, the entropy flux, J,, may be defined as the entropy flow rate per unit cross-section area. At I the entropy flow rate will be

and the entropy flux is

The heat flow rate JQ is defined as

JQ = -

A

Hence

JQ Js = - T

The rate of production of entropy per unit volume, 8, is

i.e.

from eqn (1 6.16) e = - dJS

dl

Substituting from eqn (16.18) for J, gives

(16.15)

(16.16)

(16.17)

(16.18)

( 16.19)

(1 6.20)

Equation (16.20) is the rate of production of entropy per unit volume at point I in the rod where the temperature is T. This equation was derived for thermal conduction only.

If the rod was at a uniform temperature, Le. in equilibrium, dT/dl= 0 and 0 = 0. For conduction dT/dl< 0, thus 8 is positive, i.e. entropy is produced not dissipated.

A similar calculation may be made for the flow of electricity along a wire. The wire may be considered to be in contact along its length with a reservoir at temperature T. If an electric current density, J, (=Z/A), flows due to a potential difference de, and since the wire is at constant temperature T because of its contact with the reservoir, then the electrical work must be equal to the heat transferred from the wire.

Page 338: Advanced Thermodynamics for Engineers

322 Irreversible thermodynamics

The rate of doing electrical work (power) is -J,A de, and the total rate of heat

Q = -JIA dE (16.21)

production is Q. Hence

The total change of entropy in volume dV is

J IA de

T 8dV=--

Hence, the rate of production of entropy per unit volume is, because dV=A dl,

(16.22)

(16.23)

Equation (16.23) applies for electrical flow with no heat conduction.

16.7 Thermoelectricity - the application of irreversible thermodynamics to a thermocouple

16.7.1 THERMOELECTRIC PHENOMENA

As an aid to understanding the analysis which follows, a summary will be given of the effects involved in a thermocouple. A thermocouple consists of two junctions of dissimilar metals, one held at a high temperature the other at a low temperature, as shown in Fig 16.2.

Hot junction Cold iunction

Motor

Fig. 16.2 Schematic of a thermocouple being used to drive an electric motor

Two different wires A and B are joined to form a loop with two junctions, each placed in a heat reservoir. A perfectly reversible motor is inserted in wire B. If the temperatures of the two reservoirs are different then not only will heat be transported along the wires A and B but also a flow of electrical current will occur. The following phenomena may be isolated:

(i) If the entire system is kept at a uniform temperature and the motor is driven as a generator then a current will flow around the circuit. It is found that under these circumstances heating occurs at one junction and cooling at the other. This is known as the Peltier effect and the heating (or cooling) is equal to d, where x=Peltier coefficient and I is the current. Because of the resistance of the wires it is necessary

Page 339: Advanced Thermodynamics for Engineers

The application of irreversible thermodynamics to a thermocouple 323

(ii)

(iii)

to do work to drive the generator. This work is equal to the product of current and potential difference across the wire and is called Joulean heating; this is the IZR loss, which is dissipated in the reservoirs. If the reservoirs are put at different temperatures a potential difference is set up even when no electric current flows,. This is called the Seebeck effect. Fourier conduction will also occur along the wires. If both a temperature difference and electric current occur simultaneously then a third effect OCCLUS. This is called the Thomson heat and occurs in a wire carrying a current of electricity along a temperature gradient. Consider a wire in a temperature field in the absence of an electric current. At each point in the wire there will exist a temperature as shown in Fig 16.1. Now suppose an electrical current flows. It is found that a flow of heat is required to keep the wire at the same temperature as previously. This effect is additional to Joulean heating and is proportional to the temperature gradient. The heat flux is given by

dQ dT -=CJl- dl dl

(16.24)

where Q = rate of heat transfer; I = current; dT/dl = temperature gradient in direction of I; and CJ = the Thomson Coefficient.

The above laws are purely phenomenological. The various coefficients defined in them can be used in engineering practice. It is impossible to measure some of these effects in isolation. For example, if it is attempted to measure the potential difference between two points not at the same temperature, the measuring instrument and the electrical wires used to make the connections will constitute part of a thermocouple circuit. When making connections it is necessary that the measuring points are in an isothermal region. These connections will still give rise to contact and thermally induced voltages but these are usually negligibly small.

It is not possible to measure the thermoelectric characteristics of a single material, except for the Thomson effect, because they are junction effects. Hence, the Seebeck and Peltier coefficients are the properties of pairs of metals, not single metals.

16.7.2 UNCOUPLED EFFECTS IN THERMOELECTRICITY

Consider the two possible types of uncoupled flow. Let suffix 1 refer to heat flow processes and suffix 2 refer to electrical flow processes.

Heat flow

For uncoupled heat flow the rate of entropy production per unit volume is, from eqns (16.13) and (16.20),

(16.25)

where

1 dT x----- T dl

1 - (16.26)

Page 340: Advanced Thermodynamics for Engineers

324 Irreversible thermodynamics

Electrical f low

For uncoupled electrical flow, from eqns (16.13) and (16.23)

J2X2 J,X, JI de T T T2 dl

- -- - $=- -

de

dl x

2 -

16.7.3 The Onsager relations as given by eqn (16.8) may be applied, i.e.

THE COUPLED EQUATIONS OF THERMOELECTRICITY

I J 1 = LIIXI -I- L12 x2 J2 = L 2 J I + L 2 2 x2

These become, in this case, for heat flow

de

dl 12

J - - - - - L 4 , dT - T dl Q -

for electrical flow

(16.27)

(16.28)

(1 6.29)

(16.30)

From eqn (16.18), the entropy flux Js is given by J , = JQ/T, and hence eqn (16.29) may be written

(16.3 1)

It has been assumed that both the electrical and heat flow phenomena may be represented by empirical laws of the form J = LX.

At constant temperature, Ohm’s law states I = - AA de/dl, giving

(16.32)

where A is the electrical conductivity of the wire at constant temperature.

a temperature gradient, then If dT is set to zero in eqn (16.30), i.e. the electrical current is flowing in the absence of

de de

dl dl -A - = -b2 - 2 3 L,, = A (16.33)

If the entropy fiux in the wire is divided by the electrical current flowing at constant temperature, and this ratio is called the entropy of transport, S*, then

(16.34)

Page 341: Advanced Thermodynamics for Engineers

The application of irreversible thermodynamics to thermocouple 325

The entropy of transport is basically the rate at which entropy is generated per unit energy flux in an uncoupled process. It is a useful method for defining the cross-coupling terms in the coupled equations. Thus

(16.35) L12 = TL,,S* = ITS*

By Onsager’s reciprocal relation

L,, = L12 = ITS* (16.36)

Substituting for L,,, L,, and Lzl in eqns (16.29) and (16.30) gives

and

ATS* dT de

T dl dl 1 - A-

(16.37)

(16.38)

Now if there is zero current flow (Le. JI = 0), Fourier’s law of heat conduction may be applied, giving

dT JQ=-------- - -k - A dl

However, if J I = 0 then eqn (16.38) gives

de * dT - - - -s - dl dl

Substituting this value for de fdl in eqn (16.37) gives

kdT Lll dT dT dl T dl dl

J ----.=--- Q -

Hence

Ll, = (k + ATS*’)T

Substituting the coefficients into eqns (16.29) and (16.30) gives: for heat flow

dT * de Ja = -(k + ,ITS*,) - - ,ITS - dl dl

for electrical flow

* dT de dl dl

J I = - A S -- A-

The entropy flux is obtained from eqn (16.31).

(k+ATS*’) dT * dE T dl dl

J, = - - - A S -

(16.39)

(16.40)

(16.41)

(16.42)

(16.43)

(1 6.44)

(16.45)

Page 342: Advanced Thermodynamics for Engineers

326 Irreversible thermodynamics

It is possible to define another parameter, the heat of transport Q*, where Q* = (JQ/J,) . This is basically the ‘thermal energy’ which is transported by a flow of electrical energy when there is no temperature gradient, and indicates the magnitude of the off-diagonal terms in the cross-coupling matrix in eqn (16.8a).

From eqns (16.43) and (16.44), when dT = 0

(p), = (-1TS* :)/( -1 $) = TS*

Hence * Q* = TS

(1 6.46)

( 16.47)

Equation (16.47) is interesting because it retains the basic generic form relating ‘heat’, temperature and ‘entropy’, namely that entropy is evaluated by dividing a heat transfer term by a temperature term.

It is now possible to relate the equations derived above to the various physical phenomena observed in experiments. Previously, the Seebeck effect was defined as the potential difference set up in a wire due to a temperature gradient without any current flow (Le. (de/dT),,,,). From eqn (16.44), if there is zero current flow

(16.48)

Hence, S* is a measure of the magnitude of the Seebeck effect, and the value of S* for most materials is non-zero.

If the wire is kept at a constant temperature, a flow of heat (thermal energy) will occur due to the electrical potential difference. From eqn (16.43)

(16.49)

This transport of thermal energy due to an electrical field is known as the Thomson effect.

16.7.4 THE THERMOCOUPLE

A thermocouple is a device for recording temperature at a point. It can be represented diagrammatically as shown in Fig 16.3.

EEL- Wiri

Wire X Jy:& Potentiometer

eZ I C t l

-----Ae - Hot junction Wire Y Cold junction Terminals

T”

Fig. 16.3 Schematic diagram of a thermocouple

Page 343: Advanced Thermodynamics for Engineers

The application of irreversible thermodynamics to a thermocouple 327

The thermocouple consists of two wires X and Y of dissimilar metals forming a junction at a. The ends b and c of the wires are immersed in an ice bath to form the cold junction and leads of material Z are connected to materials X and Y at points b and c, respectively, and these connections are inserted into a cold junction. These leads of material Z are then connected to a potentiometer or DVM at d and e.

The voltage representing temperature TH is obtained due to the thermoelectric effect in the leads when zero current flows, i.e. the potentiometer is balanced and J, = 0. At zero current, as previously discussed, the emf at the potentiometer (or digital voltmeter [DVM]) equals the thermal emf generated by the thermocouple.

From eqn (16.48) the potential difference between each end of a wire of material M is related to the temperature difference between the points by the equation.

(16.50) * dE),,,o = - S M dT)j,,o

where SL is the value of S* for material M . Equation (16.50) can be applied to each wire in the system shown above, to give the

following:

wire ec

EC - E, = - [: S*, dT

wire ca TH - E, = -1, S*, dT

wire ab r,

&b - Ea = -1, s*, dT

wire bd

Adding these equations to obtain the potential difference between points e and d gives the potential difference measured by the potentiometer or DVM:

TH = -1, (S; - S f ) dT

This can be written as

= E~ - E, = (SE - S:) dT 1;

(16.5 1)

(16.52)

Thus, the emf of a thermocouple at any particular temperature TH is dependent only on the materials between the hot and cold junctions if the leads bd and ce are both made of the same material. Since S* is defined in terms of J , and J , it is independent of the length of the constituent wires, and hence the emf generated is also independent of the length of the wires.

Page 344: Advanced Thermodynamics for Engineers

328 Irreversible thermodynamics

Thermocouples are normally bought as pairs of wires which have been calibrated by experimental tests. For low temperatures (up to 15OOC) Cu-Ni thermocouples may be used, for high temperatures platinum-rhodium is used. The calibration of the thermocouple is dependent only on the wires used to make the junctions; the length of thermocouples is not a parameter in the calibration, see eqns (16.50) and (16.52).

Thermocouple with junctions at T, and TH and connections in one of the measuring wires

Figure 16.3 showed an idealised thermocouple in which the connections between the thermoelectric pair of wires and those connecting the thermocouple to the measuring device were maintained at the cold junction temperature. It might be inconvenient to set up the thermocouple in this way, and a more convenient arrangement is shown in Fig 16.4.

Hot junction Cold junction

reservoir reservoir

Wire Y Wire Y

Connections - Wire Z Wire Z

Terminals T, 1123.41 Potentiometer

I J D:M

Fig. 16.4 Conventional layout of a thermocouple

It can be seen from Fig 16.4 that there are now four junctions between dissimilar materials, and each of these has the capability of generating an emf. This system will be analysed to determine the potential difference at the DVM.

Wire dc

Applying d&)J,=o = -S* dT),/,, to the system gives the following results:

E, - &d = - IT:' S*, dT

Wire cb

TC Eb - E, = -jTR s*, dT

Wire ba

Page 345: Advanced Thermodynamics for Engineers

The application of irreversible thermodynamics to a thermocouple 329

Wire ae

& e o - & = - IT: S*, dT

Wire ef

The emf at the potentiometer is given by

TR ef- ed = -ITR S*, dT - I" S*, dT - IT" S*, dT - IT' S*, dT - 1"'s: dT

TH TC TR TR

(16.53)

This is the same result given before in eqn (16.51). There is one important point which is apparent from this result and that is that the two junctions between wires Y and Z, points c and e , must be at the same temperature, i.e. (TRs)c = (TRs)e. If these junctions are not at the same temperature then each of them will act as another junction and generate an electrical potential. Hence, care must be taken when setting up a thermocouple wired in a manner similar to Fig 16.4 to ensure that points c and e are close together and at the same temperature.

16.7.5 OTHER EFFECTS IN THERMOCOUPLES

Peltier effect or Peltier heating

As defined in section 16.7.1, this effect occurs when the entire system is kept at uniform temperature and a current flows around the circuit. The temperature of the system tends to rise due to both Ohmic and Peltier heating. If the Ohmic heating is neglected initially then the Peltier coefficient, defined below in eqn (16.54), can be evaluated

JQ ~ e i ~ e r = JI (16.54)

where n = Peltier coefficient.

Fig. 16.5 Peltier heating at a junction

Page 346: Advanced Thermodynamics for Engineers

330 Irreversible thermodynamics

The Peltier heating effect at the junction in Fig 16.5 is given by

1; = (JQ>X - ( J Q > Y (16.55)

where J; is the heat that must be transferred to a reservoir to maintain the temperature of the junction at T . From eqn (16.54)

J; = X X . Y J I (16.56)

Hence, the heat transfer from the junction to maintain the temperature constant is

n X , Y J l = <JQ>X - (JQ>Y (16.57)

This is the definition of the Peltier effect, and hence the Peltier coefficient is given by

n x . y = ( h ) JI x -(:)r (16.58)

The ratio (JQ/JI)T-constant was defined previously and termed heat of transport, Q*. It was shown (eqn (16.47)) that Q* = TS*. Thus

(16.59)

It is apparent from the form of eqn (16.59) that the Peltier heating is reversible, i.e. if

* * nx,y = TSE - TS; = T ( S x - S,)

the current is reversed the direction of heat flow will be reversed.

Thomson effect or Thomson heating

This was defined previously in eqn (16.24). Rewriting this equation in the nomenclature now employed and applying it to the small element of wire, Al , shown in Fig 16.6,

( J Q ) , = oJI AT (16.60)

A!

(J&+AT J, -- -4 J'Q

Fig. 16.6 Small element of a wire subject to current flow

Assume that initially there is no current flow, but that the heat flow gives rise to a temperature drop AT. Now if an electric current is switched on it is found that a flow of heat is required to keep the wire at the same t e m p e m s as previously; this is in addition to Joulean heating (see section 16.7.1 (E)). Let the required heat transfer, to a series of reservoirs, be Jh. Under these conditions the heat flow due to heat conduction is the same as previously.

Then

(16.61) Heat flow ;ate, dQ/dr. Heat flow rate. Ohmicheating

a iT+AT dpldr. at T

Page 347: Advanced Thermodynamics for Engineers

The application of irreversible thermodynamics to a thermocouple 33 1

At any given temperature, T, the heat of transport, Q*, is (eqn (16.47))

Q* =(:)T=TS*

and hence

(JQ) = JI TS*

At a temperature, T +AT

if multiples of small terms are negected. Now, for the element

de dl

AE = -- AI

and

dT dl

AT= -- A1

Equation (16.38) can be rearranged to give

de JI * dT s - dl il dl - = - - -

Substituting the above relations into eqn (16.61) gives

2 A1 AT + JI - = JIT- dT A ds* - -

Thornson heat e x m d (0 maintain temperature

Ohmic heat generated

Now the Thomson heat was defined in eqn (16.60) as

Jn = UJ, AT

Thus

(16.62)

(16.63)

(16.64)

(16.65)

(16.66)

(16.67)

(16.68) dT

Page 348: Advanced Thermodynamics for Engineers

332 Irreversible thermodynamics

For the thermocouple shown in Fig 16.3, the difference in Thomson coefficients for the two wires is

d * * G, - GY = -T - (SX - S y )

dT (16.69)

16.7.6 SUMMARY The equations for thermocouple phenomena for materials X and Y acting between temperature limits of TH and T, are as follows:

Seebeck effect Seebeck coefficient:

Peltier effect Peltier coefficient:

nx,y= T(S,*-S;)

Thomson effect Thomson coefficient:

16.8 Diffusion and heat transfer

16.8.1 BASIC PHENOMENA INVOLVED

The classical law of diffusion is due to Fick (1856). This states that the diffusion rate is proportional to the concentration gradient, and is the mass transfer analogy of the thermal conduction law. The constant of proportionality in Fick’s law is called the diffusion coefficient. It was found from experimental evidence that the diffusion coefficient tended to vary with conditions and that better proportionality could be obtained if the rate of diffusion was related to the gradient of chemical potential; this law is due to Hartley. Figure 16.7 shows an adiabatic system made up of two parts connected by a porous membrane, or a pipe with a bore that is small compared with the mean free path of the molecules.

The Soret effect

This is a thermal diffusion effect. It is characterised by the setting up of a concentration gradient as a result of a temperature gradient.

The Dufour effect

This is the inverse phenomena to the Soret effect and is the non-uniformity of temperature encountered due to concentration gradients.

Page 349: Advanced Thermodynamics for Engineers

Diffusion and heat transfer 333

I Adiabatic wail 1

,/ / membrane Porous \

Fig. 16.7 Schematic diagram of two containers connected by a porous membrane, or small bore pipe

16.8.2 To choose the forces and fluxes it is necessary to consider the rate of entropy generation (see section 16.5, and de Groot (1951)). Suppose that a system comprising two parts, I and II connected by a hole, is enclosed in a reservoir. It will be assumed that both parts are the same volume, V , and when in thermostatic equilibrium the energy, U, and mass, m, in each part is equal, and there is an entropy, S, associated with each part of the system. (U and m were chosen as parameters because these obey conservation laws.)

The system is isolated, hence variations AU and Am in part I give rise to variations -AU and -Am in part 11. The variation in entropy due to these changes may be found from Taylor’s series:

DEFINING THE FORCES AND FLUXES

+ higher order terms (16.70)

AS11 may be found by a similar expansion, but the linear terms in AU and Am are negative. Thus, the change of entropy of the Universe is

AS = ASl + ASn = AU2 + 2 a2s AU Am + ( $)u Am2 (au am) + higher order terms . . . (16.71)

The time rate of change of AS, i.e. the rate of generation of entropy, is given by

- d (AS) = (r;) -_ (2 AU)AU + 2 ( ___ az:m)Am A 0 + 2 ( & ) A U h + ( g ) ( 2 Am) Arfz dt

=2A~[(s)AU+(*)Am}+2h[(~)Am+(~)AU] au2 au am am2 au am

(16.72)

Page 350: Advanced Thermodynamics for Engineers

334 Irreversible thermodynamics

and the rate of generation of entropy per unit volume is

d 1 d

dt 2 dt - (AS) = - - (AS) = AU (16.73)

because the combined volume of systems I and II is 2V, and the original terms were defined in relation to a single part of the system.

It was stated in eqn (16.12), from Onsager's relationship, that for a two-component system

TO = J , X i + J 2 X 2

which may be written, for this system, as

d

dr TO = - (AS) = JUXU + JmXm (16.74)

Comparison of eqns (16.73) and (16.74) enables the forces and fluxes to be defined, giving

Ju = AU Xu = A($)m

J m = h X m = A ( E ) ,

From the First and Second Laws, for a system at constant pressure,

T ds =du + p dv - p dm Hence

Substituting for

A p = v Ap - s AT and

h = p + TS = u +pv gives

Ap AT AT AT AP AT = -V - + S - + h - - Ts - = -V - + h - A($)u T T T2 T2 T T'

Hence substituting these values in eqn (16.8),

J " = L i , X u + L , , X r n J r n = L , , x ~ +L22Xrn

(16.75)

(16.76)

(16.77)

(16.78)

(16.79)

(16.80)

(16.81)

Page 351: Advanced Thermodynamics for Engineers

D i e i o n and heat transfer 335

gives

and

(16.82)

(16.83)

If the flow of internal energy which occurs due to the diffusion is defined as ( J u / J m ) A T = o = u* then

i.e. * L,,= L,U = L,,

In the stationary state, J , = 0, i.e.

h* L, h - U *

h - -

AT vT vT

(16.84)

(16.85)

(16.86)

Now, the term h - rl" may be defined as - Q*, and the significance of this is that it drives the system to generate a pressure gradient due to the temperature gradient. Hence, if two systems are connected together with a concentration (or more correctly, a chemical potential) gradient then a pressure difference will be established between the systems until h = U*.

The pressure difference is defined as

AP Q* - = -- AT vT

The significance of this result will be returned to later.

(16.87)

16.8.3

It is also possible to derive the result in eqn (16.87) in a more easily understandable manner, if the fluxes and forces defined above are simply accepted without proof.

THE UNCOUPLED EQUATIONS OF DIFFUSION

Page 352: Advanced Thermodynamics for Engineers

336 Irreversible thermodynamics

Diffusion

For the uncoupled process of diffusion, it was shown in eqn (16.78) that the thermo- dynamic force for component i, in one-dimensional diffusion, is

where pi = chemical potential. Evaluating the differential in eqn (16.88) gives, as shown in eqn (16.78)

(16.88)

(1 6.89)

Equation (16.89) applies for a single component only, but it is possible to derive the equations for multicomponent mixtures.

If Onsager’s rule for entropy flux is applied to this situation

T e = C J X

Thus

ap J, aT ax T ax TO= -Jm- + - p -

Hence the entropy production per unit volume, 6, is given as

J, ap J, aT ~2 ax T ax

e = + - p - - - Duetoisothermal Ductotcmpcnhlrc

diffusion gradient

(16.90)

(16.91)

Coupled diffusion and heat processes

Such processes are also referred to as heat and mass transfer processes. Consider the system shown in Fig 16.7. The equations governing the flow processes are:

heat

(16.92)

diffusion (mass transfer)

Jm =L2, Xt + L22X2 (16.93)

The thermodynamic forces have been previously defined. For the heat transfer process, XI = - l/T(dT/dx) (see section 16.7.2). For the mass transfer process, X, = -T a ( p / T ) / a x (see section 16.8.2).

Page 353: Advanced Thermodynamics for Engineers

Diffusion and heat transfer 337

Thus, from eqn (16.92)

and, from eqn (16.93)

(16.94)

(16.95)

Consider the steady state, i.e. when there is no flow between the two vessels and J , = 0. The effect on J , can be evaluated from eqn (16.95)

or

The specific enthalpy, h, can be related to this parameter by

Rearranging eqn (1 6.99) gives

h

The term

&P/T) 1 & ( T I T = ($1;

because T = constant, and from the definition of chemical potential, p, the term

(16.96)

(16.97)

(16.98)

(16.99)

(16.100)

dp = dg = Vdp + s dT (16.101)

Thus

(1 6.102)

Page 354: Advanced Thermodynamics for Engineers

338 Irreversible thermodynamics

Substituting for (y) P

and

in eqn (16.98) gives

h V

T2 T d(p/T) = - - dT + - dp

Hence from eqns (1 6.97) and (16.103)

VdP bl dT T2 T L 2 2 T 2

- -- - dT+-- h --

giving

If both vessels were at the same temperature then dT/dx = 0 and

Thus, as shown in eqn (16.84)

( 16.103)

(1 6.104)

(1 6.105)

( 16.106)

( 16.107)

(16.108)

The ratio (JQ/J ,JT is the energy transported when there is no heat flow through thermal conduction. Also from Onsager’s reciprocal relation (JQ/J,JT = L21/L22. If this ratio is denoted by the symbol U* then eqn (16.105) can be written (eqn (16.86))

dp ( h - U * ) dT vT - - -

Now the difference h - U* was denoted - Q*, which is called the heat of transport. Using this definition, eqn (16.86) may be rewritten (eqn (16.87))

dp Q* - = - dT VT

Page 355: Advanced Thermodynamics for Engineers

Diffusion and heat transfer 339

The signficance of Q*

Q* is the heat transported from region I to region 11 by diffusion of the fluid. It is a measure of the difference in energy associated with the gradient of chemical potential. In a simple case it is possible to evaluate Q*. Weber showed by kinetic theory that when a gas passes from a vessel into a porous plate it has a decrease in the energy it carries. The energy carried by the gas molecule in molecular flow through the passages in the porous medium is smaller by RT/2 than it was when the motion was random.

Hence Q* = - RT/2 for flow through a porous plug. The energy, Q*, is liberated when the molecules enter the plug and the same amount is absorbed when the molecules emerge from the plug. When Q* is dependent on temperature, as in the thermal effusion case, it would ap ar reasonable that if Q* is liberated on entering the plug at temperature T then Q*+dQ would be absorbed on the molecule leaving the plug, when the eneral temperature is T + dT. This is erroneous and a qualitative explanation follows. Q is the amount by which the mean energy per unit mass of molecules that are in the process of transit through the plug exceeds the mean energy of the bulk of the fluid. Hence, if Q* were not applicable to both sides of the plug then the principle of conservation of energy would be contravened. The difference of average values of the bulk fluid energy on either side of the plug has been taken into account by the different values of the enthalpy, h , on each side of the plug.

F! $

Now, if

Q* = -RT/2 (16.109)

and for a perfect gas

pv = RT (1 6.1 10)

then

dp Q* RT R p

dT vT 2vT 2v 2T

Integrating across the plug from side I to side 11,

- - _ -

or

(16.111)

(16.112)

(16.113)

This is called the Knudsen equation for molecular flow. From eqn (16.87) over an increment of distance, Ax

Q* A p = -- AT vT

This means that if Q* is negative, as in the thermal effusion case, the pressure increment is of the same sign as the temperature increment and molecules flow from the cold side to the hot side.

Page 356: Advanced Thermodynamics for Engineers

340 Irreversible thermodynamics

16.8.4 THERMAL TWSPIRATION

It will be shown in this section, without recourse to Onsager’s relationship, that L,, = L2,. This derivation will be based on statistical mechanics, and this is the manner by which Onsager derived his relationship. The problem considered is similar to that discussed in section 16.8.3, when flow through a porous plug was examined. In this case the flow will be assumed to occur along a small bore pipe in which the cross-sectional area is small compared with the mean free path of the molecules; this is what happens when gas flows through the pores in a porous membrane.

It can be shown from statistical thermodynamics that the number of particles striking a unit area in unit time is

P - 4 A (2nkTm)’I2 _ - (1 6.1 14)

where ri, = number of particles striking wall A = area of wall p = ‘pressure’ of gas k = Boltzmann’s constant T = temperature m = mass of ‘particle’

Consider a hole of area A , then ri, would be the number of particles passing through this hole. These particles would carry with them kinetic energy. The hole also makes a selection of the atoms which will pass through it and favours those of a higher energy because the frequency of higher velocity particles hitting the wall is higher. It can be shown that the energy passing through the hole is given by

E A +g2 Equations (16.114) and (16.115) may be written in the form

_ -

E‘ A ( NkT )( 2m kT) l l z ( : kT)(-!!?) 3n _. -

where N = Avogadro’s number.

The groups of terms can be defined in the following way:

(&) = density

(E) = velocity

(16.115)

(1 6.1 16)

( 16.1 17)

3 - kT = energy 2

Page 357: Advanced Thermodynamics for Engineers

Diffusion and heat transfer 341

Nm'/2/n'/2 and 4N/3n1" are weighting factors. Thus

ri, - = density x mean velocity A

E - = density x mean velocity x mean energy A

The weighting factors show that the particles passing through area A carry with them an unrepresentative sample of the energy of the system from which they come.

This example can be expanded into an apparatus that contains two systems separated by a porous wall. It will be assumed that the pores in the wall are much shorter than the mean free path of the molecules. The system is shown in Fig 16.8. The gases on either side of the partition are the same, but one side of the system is heated and the other side is cooled.

Cooling - Q

Heating + Q

Porous wall

Fig. 16.8 Arrangement for flow through a porous plug

From eqn (16.1 14) the net flux of molecules passing through the wall is given by

and the net energy passing through the wall is given from eqn (16.1 15) as

(16.1 18)

(16.119)

It is possible to substitute for the pressure of a monatomic gas by the expression = e-ah-3(2zm)3/2p-5/2 ( 1 6.1 20)

and the temperature by

T = I/kp (16.121)

Then eqns (16.118) and (16.119) become

( 16.122)

( 16.123)

Page 358: Advanced Thermodynamics for Engineers

342 Irreversible thermodynamics

If the state denoted by the primed symbols is constant, then differentiating gives 4:) = (2~mzh-~)B-~e-"[-6j3-' dB - 2 da]

4:) = (2nr~h-~)B-~e-"[ -2 dp - /? da]

Equations (16.124) and (16.125) may be written in the form 4:) = LE,^ dB + LE,, da

and

( 1 6.124)

(16.125)

(16.126)

(16.127)

where

LE,s =

L . - - 4 ~ 7 m B - ~ h - ~ e - ~ E . " -

L. nc B = -4nmfi -3h-3e-a

L . nc a = -2nmB-2h-3e-a

Hence Lt., = LkB, which is equivalent to the result given by Onsager, that L,, = L2, and has been proved frGm Statistical thermodynamics.

16.9 Concluding remarks This chapter extended 'thermodynamics' into a truly dymanic arena. The processes involved are non-equilibrium ones which are in a steady, dynamic state. The concepts of entropy generation and coupled phenomena have been introduced, and Onsager's reciprocal relation has been used to enable the latter to be analysed.

Thermoelectric phenomena have been considered and the coupling between them has been described. The major thermoelectric effects have been defined in terms of the thermodynamics of the device. Coupled diffusion and heat transfer processes have been introduced and analysed using these techniques, and the conjugate forces and fluxes have been developed. Finally, statistical thermodynamics was used to demonstrate that Onsager's reciprocal relation can be proved from molecular considerations.

PROBLEMS

1 The emf of a copper-iron thermocouple caused by the Seebeck effect, with a cold junction at O'C, is given by

a 2 2 a3 3

2 3 & = a , t + - t + - t V

Page 359: Advanced Thermodynamics for Engineers

Problems 343

where a! = -13.403 x V/"C a2 = +0.0275 x V/("C)' aj = +0.00026 x V/(°C)3 t = temperature ("C)

If the hot junction is at t = 100"C, calculate

(a) the Seebeck emf; (b) (c) the net Thornson emf; (d)

the Peltier effect at the hot and cold junctions;

the difference between the entropy of transport of the copper and iron.

[1.116 mV; 3.66 mV; 3.00 mV; -8.053 x VI

2 The emf of a copper-iron thermocouple with its cold junction at 0°C is given by

E = -13.403t+0.0138t2+0.0001t3 pV

where t = temperature ("C).

Show that the difference in the Thomson coefficient for the two wires is

7.535 + 0.1914t + 0.0006t2 pV/"C

3 If a fluid, consisting of a single component, is contained in two containers at different temperatures, show that the difference in pressure between the two containers is given by

dp h - u *

dT vT

u*=the energy transported when there is no heat flow through thermal conduction,

v = specific volume, T = temperature.

-=-

where h = specific enthalpy of the fluid at temperature T,

4 A thermocouple is connected across a battery, and a current flows through it. The cold junction is connected to a reservoir at 0°C. When its hot junction is connected to a reservoir at 100°C the heat flux due to the Peltier effect is 2.68 mW/A, and when the hot junction is at 200°C the effect is 4.11 mW/A. If the emf of the thermocouple due to the Seebeck effect is given by E = at + bt2, calculate the values of the constants a and b. If the thermocouple is used to measure the temperature effect based on the Seebeck effect, i.e. there is no current flow, calculate the voltages at 300°C and 200°C.

[5.679 x mV/K; 7.526 x mV/(K)2; 0.6432 mV; 1.437 mV]

5 A pure monatomic perfect gas with cp = 5%/2 flows from one reservoir to another through a porous plug. The heat of transport of the gas through the plug is -%T/2. If the system is adiabatic, and the thermal conductivities of the gas and the plug are negligible, evaluate the temperature of the plug if the upstream temperature is 60°C.

[73"C]

Page 360: Advanced Thermodynamics for Engineers

344 Irreversible thermodynamics

A thermal conductor with constant thermal and electrical conductivities, k and 1 respectively, connects two reservoirs at different temperatures and also carries an electrical current of density, J,. Show that the temperature distribution for one- dimensional flows is given by

d2T Jla dT J: dx2 k d x A

+ - = o ----

where (J is the Thomson coefficient of the wire.

A thermal conductor of constant cross-sectional area connects two reservoirs which are both maintained at the same temperature, To. An electric current is passed through the conductor, and heats it due to Joulean heating and the Thomson effect. Show that if the thermal and electrical conductivities, k and A, and the Thomson coefficient, (I, are constant, the temperature in the conductor is given by

Show that the maximum temperature is achieved at a distance

k(eJruL/k - 1)

(:)=&h[ JlaL } Evaluate where the maximum temperature will occur if J,aL/k = 1, and explain why it is not in the centre of the bar. Show that the maximum temperature achieved by Joulean heating alone is in the centre of the conductor.

[x/L = 0.541 ]

Page 361: Advanced Thermodynamics for Engineers

Fuel Cells

Engineering thermodynamics concentrates on the production of work through cyclic devices, e.g. the power to drive a vehicle as produced by a reciprocating engine; the production of electricity by means of a steam turbine. As shown previously, these devices are all based on converting part of the Gibbs function of the fuel to useful work, and if an engine is used some energy must be ‘thrown away’: all engines are limited by the Carnot, or Second Law, efficiency.

However, there are some devices which are capable of converting the Gibbs energy of the fuel directly to electricity (a form of work); these are called fuel cells. The advantage of a fuel cell is that it is not a heat engine, and it is not limited by the Carnot efficiency. The thermodynamics of fuel cells will be developed below.

The concept of the fuel cell arises directly from the operating principle of the electric cell, e.g. the Daniel1 cell, and as early as 1880 Wilhelm Ostwald wrote: ‘I do not know whether all of us realise fully what an imperfect thing is the most essential source of power which we are using in our highly developed engineering - the steam engine.’ He realised that chemical processes could approach efficiencies of 100% in galvanic cells, and these were not limited by the Carnot efficiency. Conversion efficiencies as high as 60-80% have been achieved for fuel cells, whereas the practical limit even for sophisticated rotating machinery is not much above 50%. A further benefit of the fuel cell is that its efficiency is not reduced by part load operation, as is the case for all heat engines. Hence, if a fuel cell operating on hydrocarbon fuels can be achieved it will improve ‘thermal efficiency’ significantly and reduce pollution by CO, and NOx. The current situation is that successful commercial fuel cells are still some way from general use but small ones have been used in specialist applications (e.g. space craft) and large ones are being developed (e.g. 1 MW by Tokyo Gas, Japan). Figure 17.1 shows a proposal for a fuel cell to power a motor vehicle using methanol (CH,OH) as its primary fuel. In this case the methanol is used to generate hydrogen for use in a hydrogen-oxygen fuel cell. Figure 17.l(a) shows the hydrogen generator which is based on the water gas reaction, and Fig 17.l(b) depicts the fuel cell itself. This arrangement has the advantage that the fuel can be carried in its liquid phase at atmospheric conditions, which is much more convenient than carrying hydrogen either in gas, liquid or hydride form.

The theory of fuel cells can be developed from the previously derived thermodynamic principles, and it shows how equilibrium reversible thermodynamics can be interwoven with irreversible thermodynamics. Before developing the theory of the fuel cell itself it is

Page 362: Advanced Thermodynamics for Engineers

346 Fuel cells

Fig. 17.1 Proposed hydrogen fuel cell for vehicle applications

necessary to consider simpler electrical cells. A good place to start is the Daniell cell which produces electricity through consuming one electrode into solution and depositing onto the other one.

17.1 Electric cells

A schematic of a Daniell cell is shown in Fig 17.2. This can be represented by the convention

(17.1)

where the I represents an interface, or phase boundary. The convention adopted in representing cells in this way is that the electrode on the right-hand side of eqn (17.1) is positively charged relazive to that on the left if the reaction takes place spontaneously.

Zn I ZnSO, I cuso, I Cu

Page 363: Advanced Thermodynamics for Engineers

Electric cells 347

-netectrode f +n electrode zinc rod

//F I: acid

copper / sulfate solution

/ copper vessel

Fig. 17.2 Schematic diagram of a Daniell cell

To understand how the Daniell cell produces a potential difference and a current it is

(17.2)

which indicates that the zinc reacts with the copper sulfate solution to produce copper and zinc sulfate solution. The notation aq indicates an aqueous solution of the salt. If the reaction in eqn (17.2) takes place in a constant volume container using 1 kmol Zn (45 kg Zn) then 214 852 kJ of heat must be transferred from the container to maintain the temperature of the system at 25°C. This reaction is similar to a combustion reaction, and must obey the First Law of Thermodynamics

SQ =dU+SW (17.3)

necessary to consider the basic reaction involved,

Zn + CuSO,(aq)-Cu + ZnSO,(aq)

where 6 W = 0 in this case. Hence

SQ, = -214 852 kJ/kmol Zn = ucu + uznso4(oq) - uzn - ucuso4(oq) (17.4)

The reaction described above is basically an open circuit reaction for a Daniell cell, and SQ, represents the open circuit energy released. If, as is more usual, the reaction takes place as work is being taken from the cell then supposing a current I flows for time, t , then, from Faraday’s laws of electrolysis, the ratio of the amount of substance (Zn) dissolved to the valency of the element, n / z , is proportional to the electric charge passed, i.e.

n / z = Q, n / z = FQ (17.5)

When n / z = 1 kmol, Q = 9 6 485 kC (where kC denotes kcoulomb), and hence F =96 485 kC/kmol. This is known as the Faraday constant, and is the product of Avogadro’s number and the charge on a proton. Now, if the potential between the electrodes is E, then the work done is

(17.6) 6W = 2E,F V kC/kmol = 2E,F kJ/kmol

Page 364: Advanced Thermodynamics for Engineers

348 Fuel cells

In a Daniell cell the potential at zero current (i.e. on open circuit), which is called the emf, E, = 1.107 V at 25°C. If it is assumed that the cell can maintain this potential at low currents, then

6W = 2 x 1.107 x 96 485 = 213 618 kJ/kmol

If the reaction described above takes place isothermally in a closed system then it must obey the First Law, which this time is applied to the closed circuit system and gives

dU=SQ-SW =dQR- SW=dQ,-213 618 W/km01 (17.7)

Now the change in internal energy is simply due to the chemical changes taking place, and for an isothermal reaction must be equal to SQ, defined in eqn (17.4), giving

(17.8)

This indicates that heat must be transferred from the cell to maintain the temperature constant. This heat transfer is a measure of the change of entropy contained in the bonds of the product (ZnSO,) compared with the reactant (CuSO,), and

AS= -1234/298= -4.141 kJ/K (17.9) In this approach the Daniell cell has been treated as a thermodynamic system - a black

box. It is possible to develop this approach further to evaluate the electrical performance of the cell. Suppose that an amount of substance of Zn, dn, enters the solution at the negative pole, then it will carry with it a charge of zF dn, where z is the valency (or charge number of the cell reaction) of the Zn, and is 2 in this case. The Coulombic forces in the cell are such that an equal and opposite charge has to be absorbed by the copper electrode, and this is achieved by the copper ions absorbing electrons which have flowed around the external circuit. The maximum work that can be done is achieved if the cell is reversible, and the potential is equal to the open circuit potential; thus

6W= zFE dn (1 7.10)

dQR= -214 852 + 213 618 = -1234 W/km01

However, the total work that could be obtained from a cell if it changed volume would be 6 W = zFE dn + p dV (17.11)

thus, applying the First Law, and assuming the processes are reversible gives d U = d Q - 6 W = T d S - p d V - z F E d n (17.12)

and hence the electrical work output is -zFE dn = d U + p dV- T d s = d G = G2 - G, (1 7.13)

For a cell which is spontaneously discharging, G2 < G, , and hence dW= -dG (17.14)

The equations derived above define the operation of the Daniell cell from a macroscopic viewpoint. It is instructive to examine the processes which occur at the three interfaces shown in eqn (17.1). Hence

(1 7.15) I Zn - zn" + 2e at zinc electrode Zn" + CuSO, - ZnSQ, + Cu++ in solution Cu++ + 2e - Cu at copper electrode

Page 365: Advanced Thermodynamics for Engineers

Electric cells 349

This means that the zinc is ‘dissolved’ by the sulfuric acid at the zinc electrode and a zinc anion enters the solution. Meanwhile, two electrons are left on the zinc electrode (because the valency of zinc is 2) and these are free to travel around the circuit, but cause the zinc electrode to be at a negative potential, i.e. it is the cathode. The zinc anion reacts with the copper sulfate in solution to form zinc sulfate and release a copper ion which migrates to the copper electrode, where it withdraws electrons from the electrode giving it a positive potential. Hence, the Daniell cell consists of electrons (negative charges) travelling around the outer circuit, from the cathode to the anode, while positive ions travel through the solution from cathode to anode. (Note: the convention for positive electric current is in the opposite direction to the electron flow; the current is said to flow from the anode to the cathode.) The net effect is to maintain the potential difference between the electrodes constant for any given current: this is a state of dynamic equilibrium. It can be seen that the electrochemical cell is a situation governed by thermodynamic equilibrium and steady state (irreversible) thermodynamics (see Chapter 16).

The reactions defined in eqn (17.15) resulted in electrons flowing from the zinc to the copper (this would be defined as a current flowing from the copper [anode] to the zinc [cathode]), and the potential on the anode would be higher than the cathode. If the cell were connected to a potential source (e.g. a battery charger), such that the potential difference of the source were slightly higher than the cell emf, then the current flow could be reversed and the reaction would become

(17.16) I Cu - Cu++ + 2e at Cu electrode

Cu++ + ZnSO, - Zn++ + CuSO, in solution

a++ + 2e -. Zn at Zn electrode

Hence, the Daniell cell is reversible if the current drawn from (or fed to) it is small. The Daniell cell can be used to ‘generate’ electricity, by consuming an electrode, or to store electricity.

Although the Daniell cell was one of the early examples of a device for generating electricity, it is relatively difficult to analyse thermodynamically because it has electrodes of different materials. A simpler device will be considered below to develop the equations defining the operation of such cells, but first it is necessary to introduce another property.

17.1.1 ELECTROCHEMICAL. POTENTIAL.

To be able to analyse the cell in more detail it is necessary to introduce another parameter, the electrochemical potential, p, of ions. This is related to the chemical potential by

(17.17) pi = p i + z i V F where

zi = valency of (or charge on) the ion (+/ - ); V = inner electric potential of the phase; F = Faraday constant (96 485 kJ/kmol).

It was previously stated that the chemical potential was a ‘chemical pressure’ to bring about a change of composition to achieve a state of chemical equilibrium. In a similar

Page 366: Advanced Thermodynamics for Engineers

350 Fuel cells

manner the electrochemical potential is an 'electrical pressure' that pushes the system towards electrical equilibrium. The electrochemical potential also contains a term for the change in capacity to do work through electrical processes, introduced to the system by the introduction of a unit of mass.

The value of w is the potential of the phase, and is obtained from electrostatic theory. It is the work done in bringing a unit positive charge from infinity to within the phase. Hence, p is the sum of the chemical potential, which has been based on the 'chemistry' of the material, and an electrical potential. It defines the energy that can be obtained from mechanical, chemical and electrical processes.

17.1.2 THERMODYNAMIC ORIGIN OF EMF

Consider a simple galvanic cell, such as that shown in Fig 17.3. This consists of two electrodes, A and B, made of different materials, an electrolyte, and two connecting wires of identical material. The cell can be represented as

(1 7.18) Cu' \%A I,,electrolyte IX4B Ix3Cu"

The junctions of interest here are X, and X4; junctions X, and X3 are potentially thermocouples (because they are junctions of two dissimilar metals) but these effects will be neglected here.

Fig. 17.3 Simple galvanic cell

Considering the electrochemical potential between points 1 and 2

P1-P2=PI -P6- @ S - & ) - @4-p5)- @3-P4)- @ 2 - L 1 3 )

= P C u ' - P A - @ e I e c - f i A ) - @elec -Pclcc) - @ B -Pclec) - @CU' - P B )

= P c u ' - P c u "

= P c u , + Z c u , W c u J - ( P C U , + Z C u " W C u ' F )

= Z C " F ( WC" . - WCU, ) (1 7.19)

The difference in chemical potential of the two electrodes (pcus - pas) is zero because both electrodes are of the same composition and are at the same temperature. Equation

Page 367: Advanced Thermodynamics for Engineers

Fuel cells 351

(17.19) defines the maximum work that can be obtained from the cell by a reversible process, i.e. a small flow of current. The difference, - qcu”, is equivalent to the open circuit potential, or emf, E , and hence

(17.20)

17.2 Fuel cells The above sections have considered how electrical work can be obtained by the transfer of material from one electrode to another. In the Daniell cell the Zn electrode is consumed, and ultimately the cell would cease to function. The situation in the Daniell cell can be reversed by recharging.

If it is desired to manufacture a cell which can operate without consuming the electrodes then it is necessary to supply the fuel along an electrode: such a device is called a fuel cell. An example of a fuel cell is one which takes hydrogen (H,) and combines it with chlorine (Cl,) to form hydrochloric acid (HC1).

A practical cell similar to that in Fig 17.3 can be constructed in the following way:

-CU’ I Pt 1 H,(g;p) I HCl(aq; C ) I Cl2(g; P ) I Pt I CU” +

where (g; p ) denotes a gas at pressure, p and (a4; c) denotes an aqueous solution with concentration, c

(17.21)

This is basically a cell using platinum electrodes (Pt) in an electrolyte of dilute hydrochloric acid (HCI). The platinum acts both as the electrodes and a catalyst, and when it is dipped into the dilute hydrochloric acid the system acts like an electric cell and hydrogen is evolved at the left-hand wire and chlorine at the right-hand one; electricity flows around the circuit and the energy is derived by breaking the bonds in the HC1 molecules.

If the cell is to operate as a fuel cell then it will not break down the hydrochloric acid, as described above, but it will have hydrogen and chlorine delivered as fuels, which will be combined in the fuel cell to produce hydrochloric acid and electrical power. The operation is then defined by the reaction

H,(g;p) + Cl,(g;p)-2HCl(aq; c ) (17.22) where p is the pressure of the gases in the cell. The reactions at the interfaces are of interest because they produce the electrons and govern the potential. At the left-hand electrode, if the process is considered to be reversible

1 - H,(g) + Pt + a4 - Pt(e) + H’(aq) 2

(17.23)

This reaction indicates that for hydrogen gas (H,) to be evolved at the electrode a hydrogen ion (H’) must be released into the solution and an electron must be left on the electrode. Now, it is not obvious from eqn (17.23) which direction the reaction will go; this is governed by considering the pair of processes occurring at the platinum electrodes. At the right-hand electrode

1 - Cl,(g) + Pt + aq - Pt’ + Cl-(aq) 2

(17.24)

Page 368: Advanced Thermodynamics for Engineers

352 Fuel cells

Equations (17.23) and (17.24) in combination show that for the reaction shown in eqn (17.22) the electron current must flow from the Pt-H, electrode to the R-C1, electrode. Hence, the Pt-H, is a cathode and the Pt-C1, electrode is an anode. It is apparent that when the cell is operating open circuit it will build up a potential difference between the electrodes. However, once a current starts to flow, that potential difference will be affected by the ability of the H' and C1- ions to undertake reaction (17.22): this then introduces irreversible thermodynamics.

First, consider the device operating open circuit. It is in a state of equilibrium and AG)p,T=O for an infinitesimal process. If it is assumed that the equilibrium can be perturbed by an amount of substance 1/2 dn of hydrogen and chlorine being consumed at the electrodes, then these will generate dn H' and C1- in the electrolyte and transfer dn F of charge from Cur to Cu". Thus

Cu" (e)-Pt(e) (17.25)

Pt(e)-Cu' (e) (17.26)

The reactions taking place in the cell can now be examined to give

(17.27)

The change of molar Gibbs function across the cell is

AG = AGl + AG, + AGj + AG4 ( 1 7.28)

which gives

(17.29)

This equation consists of a mixhxe of chemical potential ( p ) and electrochemical potential @) terms. Expanding the latter gives

&, = PH& + FV,,

PCG, = P C & + FVWh

P c u , ( e ) = P C ~ , + F V c u ,

P c u . ( e ) = P C ~ ' + F V c u '

These terms may be substituted into eqn (17.29) to give

(17.30)

(17.3 1)

Page 369: Advanced Thermodynamics for Engineers

Fuel cells 353

Now, at equilibrium AG),,T = 0 and F(q!J,,. - q!JCum) = FE, thus (introducing the valency, z , to maintain generality)

(17.32)

where pr = p/po and a = the activity coefficient of the particular phase. The activity coefficient, a , can be defined in such a way that the chemical potential ofa solution is

(17.33)

In eqn (17.33) p,* is a function of temperature and pressure alone, and a, takes account of the interaction between the components in the solution. It can be shown that at low concentrations the activity coefficients are approximately equal to unity. If the standard chemical potential terms (pa) at temperature T are denoted AGF then

p, =,ut* + %T In a,x ,

(17.34)

where the term -AG:/zF is called the standard emf of the cell, Eo. Equations (17.32) and (17.34) show quite distinctly that the processes taking place in

the fuel cell are governed by similar equations to those describing combustion. The equations are similar to those for dissociation, which also obeys the law of mass action. In the case of gaseous components the activity can be related to the partial pressures by a, = pr/po, where pi = partial pressure of component i.

17.2.1 EXAMPLE: A HYDROGEN-OXYGEN FUEL CELL

Consider a fuel cell in which the fuel is hydrogen and oxygen, and these gases are supplied down the electrodes. A suitable electrolyte for this cell is aqueous potassium hydroxide (KOH). The cell can be defined as

The basic reactions talung place at the electrodes are

H, + 20H- - 2H,O + 2e at hydrogen electrode

at oxygen electrode 1 - 0, + H,O + 2e - 20H- 2

and these can be combined to give the overall reaction

1

2 H2 + - 0 2 + H2O

(17.36)

(17.37)

It can be seen that the potassium hydroxide (KOH) does not take part in the reaction, and simply acts as a medium through which the charges can flow. It is apparent from eqn

Page 370: Advanced Thermodynamics for Engineers

354 Fuel cells

(17.36) that the valency ( z ; ) of hydrogen is 2. From eqn (17.34) the cell open circuit potential, the standard emf, is given by

0 A G 9 8 -(-30 434.3 - 0.5 x 52 477.3 - [-239 081.7 - 46 370.11) E =--=- Z H , F 2 x 96 485

= 1.185 v (17.38)

The standard emf in eqn (17.38) has been evaluated by assuming that the partial pressures of the gases are all 1 atmosphere. In an actual cell at equilibrium they will be controlled by the chemical equation, and the actual emf achievable will be given by an equation similar to eqn (17.34) which includes the partial pressures of the constituents,

\,,,, / z;F z;F \;;,, /

Equation (17.39) can be written in a shorter form as

(17.39)

(17.39a)

where the stoichiometric coefficients, v, are defined as positive for products and negative for reactants. Hence the emf of the hydrogen-oxygen fuel cell is

(17.40)

Considering again the term for Eo, it is directly related to the equilibrium constant, Kpr, by

(17.41)

giving

(17.42) %T %T

Z; F E = - In K , - - In

Examination of eqn (17.42) shows that when the constituents of the cell are in equilibrium, which is defined by

the potential output of the cell is zero. Hence, like a combustion process, the fuel cell only converts chemical bonds into another form of energy when it is transferring from one equilibrium state to another. A combustion process converts the bond energy of the reactants into thermal energy which can then be used to drive a powerplant; the fuel cell converts the bond energy of the fuel directly into electrical energy, and the amount of electricity produced is not constrained by the Carnot efficiency. The Second Law does

Page 371: Advanced Thermodynamics for Engineers

Fuel cells 355

limit the energy output but only defines the equilibrium condition. A fuel cell is basically a direct conversion device which is constrained mainly by the First Law.

An interesting difference between the fuel cell and the combustion process is that many fuel cells attempt to release the higher enthalpy of reaction, whereas combustion usually releases the lower enthalpy (or internal energy) of reaction. In the case of the H2-02 fuel cell the product (H20) is in the liquid phase and not the vapour phase as is usual after combustion. However, the potential to do work is not changed by the phase of the products (the chemical potential of H,O at any particular temperature is the same in the liquid and vapour phases). This means that while fuel cells are controlled largely by the First Law, they have thermal efficiencies significantly below 100% based on the higher enthalpy of reaction: they have to transfer away the difference between the higher and lower enthalpy of reaction in the form of heat if they are going to operate isothermally.

17.2.2

The effect of the temperature of operation on the emf of the cell can be evaluated by differentiating eqn (17.42) with respect to T , in this case keeping the pressure constant. This gives

EFFECT OF TEMPERATURE ON FUEL CELL OPERATION

Now, from the Van't Hoff equation (12.101).

d Q P

dT R T 2 - (In K p , ) = -

Substituting from eqn (17.44) into eqn (17.43) gives

E ziF R T 2 T ziFT T

+ - = -

(17.43)

(17.44)

(17.45)

Equation (17.45) shows that the change of emf is related to the heat of reaction, Qp. Now, for exothermic reactions, Qp is negative, and this means that the emf of a cell based on an exothermic reaction decreases with temperature. This is in line with the effect of temperature on the degree of dissociation in a combustion process, as discussed in Chapter 12.

17.2.3 EXAMPLES

Example 1

A Pb-Hg fuel cell operates according to the following equation:

Pb + Hg,Cl, - PbCl, + 2Hg

Page 372: Advanced Thermodynamics for Engineers

356 Fuel cells

and the heat of reaction, Qp = -95 200 kJ/kmol Pb. The cell receives heat transfer of 8300 kJ/kmol Pb from the surroundings. Calculate the potential of the cell, and evaluate the electrical work produced per kilogram of reactants, assuming the following atomic weights: Pb - 207; Hg - 200; C1 - 35.

Solution

This problem can be solved by a macroscopic approach to the cell as a closed system. Applying the First Law gives

-zFEdn=dU-SQ,,+TdS-pdV = dG - SQ,,

Now this value of energy transfer has to be used in eqn (17.34) in place of AGO, giving

(AG:- SQ,,,) -95 200 - 8300 = 0.5363 V - E = - _ -

z F 2 x 96 487

The quantity of work obtained per kg of reactants is given in the following way. Assuming the potential of the cell is not reduced by drawing a current, then the charge transferred through the cell per kmol Pb is 2 F (because the valency of lead is 2). Hence the work done per kmol Pb is

SW = 2 x 96 485 x 0.5363 = 103 500 kJ/kmol Pb

giving the work per unit mass of reactants as

103 500

207 + 2 x (200 + 35) = 152.9 kJ/kg

SW

mmct -=

Example 2

A hydrogen-oxygen fuel cell operates at a constant temperature of 227°C and the hydrogen and oxygen are fed to the cell at 40 bar, and the water is taken from it at the same pressure. Evaluate the emf at this condition, and the heat transfer from the cell.

Solution

This question requires the application of eqn (17.42), giving

From tables at 500 K, K , = 7.92127 x bar-”2. Hence the open circuit potential is

E = 8.3143 x 500 { ln(7.92127 x x 11’2) - ln( 4yb’:ti/ 2 x 96 487

= 1.176V

Page 373: Advanced Thermodynamics for Engineers

Fuel cells 357

The maximum work output from the fuel cell based on eqn (17.14) is

W = -mAGO,= ri?( -239 081.7 - 86 346 - (0.5 x [-95 5441 + [-58 4721)) = h(-219.18) MJ/kmol H,

Also, at 500 K the lower enthalpy of reaction of the process is 244.02 MJ/kmol, which gives a higher enthalpy of reaction of

Qp = -(244.02 x lo3 + 18 x 1831) = -277.0 x lo3 = -277.0 MJ/kmol H,

This is equal to the change of enthalpy of the working fluid as it passes through the cell and hence applying the steady flow energy equation to the fuel cell gives

Q - w = ri?(Ah)

giving

Q = k ( A h ) + W = -277.0 + 219.18 = -57.82 MJ/kmol H,

Thus, the efficiency of the fuel cell, based on the higher enthalpy of reaction, is

219.18

277.0 q=-- - 0.7913

If it were evaluated from the lower enthalpy of reaction, which is usually used to calculate the efficiency of powerplant, then the value would be

219.18

244 v=-- - 0.8983

Example 3

The operating pressure of the fuel cell in Example 2 is changed to 80 bar. Calculate the change in emf.

Solution

If the pressure is raised to 80 bar, this only affects the pressure term because K , = f(T), and hence

E , - E , = - S T { [ h K p , - h ( 80x11”)] - [ In KpT - In ( 40xl1”) ]} z ~ F 80 x 80’/, 40 x 40‘’2

%T

z ~ F = - ln(2’/,) = 0.0075 V

giving

E, ,= 1.176+0.0075= 1.183 V

This shows that the potential of the hydrogen-oxygen fuel cell increases with pressure. This is similar to the effect of pressure on the hydrogen-oxygen combustion reaction, when dissociation decreased from 0.0673 to 0.0221 as the pressure was increased from 1 bar to 100 bar for the stoichiometric combustion of methane in Chapter 12. Hence, if the

Page 374: Advanced Thermodynamics for Engineers

358 Fuel cells

amount of substance in the reactants is not equal to the amount of substance in the products then the emf of the cell will be a function of the pressure of operation.

17.3 Efficiency of a fuel cell The fuel cell is not 100% efficient, and it is possible to define its efficiency by the following equation:

Maximum useful work output w,,

Heat of reaction Qp

=- I I = (17.46)

Now the maximum useful work output obtainable is defined by the change of Gibbs function, i.e.

W, = AG = GprodUc, - G,,,

and

Thus, the efficiency is

Qp

(1 7.47)

(17.48)

(17.49)

The last expression in eqn (17.49) has been written in terms of the calorific value of the fuel because this is usually positive, and hence the efficiency is defined by the sign of the change of entropy. If the entropy change is positive then the efficiency is greater than unity; if it is negative then it is less than unity. Consider now the hydrogen-oxygen reaction described above. If the fuel cell is maintained at the standard temperature of 25°C. and the processes are assumed to be isothermal, then

AG = ( h ? ) H 2 0 - T(SH~O - S H ~ - o-5soz) = -241 820 - 298(188.71- 130.57 -0.5 x 205.04) = -228 594.8 k.J/km~l

and Qp = -241 820 k.J/kmol. Hence the efficiency is

228 594.8

241 820 = 0.945 t l =

(17.50)

(17.5 1)

This efficiency can be evaluated using eqn (17.49) to give

298(188.71 - 130.57 - 0.5 x 205.04) = 1 - 0.0547 = 0.945

T AS q = l + - - - 1 +

Q; 241820

(17.52)

Page 375: Advanced Thermodynamics for Engineers

Thermodynamics of cells working in steady state 359

This is exactly the same value as was obtained in Chapter 2 on exergy and availabilty. The rational efficiency of this, and any, fuel cell would be 100%. The efficiency of this cell is less than 100% because the change of entropy as the reactants change to products is negative. Hence, if the value of TAS < 0 then the cell will be less than 100% efficient, and this is because ‘heat’ equivalent to TAS must be transferred with the surroundings, i.e. to the surroundings. If TAS > 0 then the efficiency can be greater than loo%, and this comes about because energy has to be transferred to the fuel cell from the surroundings to make the process isothermal.

17.4 Thermodynamics of cells working in steady state

The previous theory did not relate to a cell in which current was being continually drawn from it. Under such circumstances, concentration gradients might be set up in the cell, and it will be governed by the theories of irreversible thermodynamics (Chapter 16). It can be shown for a fuel cell that a possible force term is the gradient of electrochemical potential, which is similar to the use of the chemical potential gradient when considering diffusion processes (note that when considering thermal and mass diffusion in combination the gradient of the ratio of chemical potential and temperature ( p / T ) was used.) Hence, the force term is

(17.53)

In this analysis the flux term is taken to be the movement of charge per unit area through the cell. Since the charge is carried on the ions travelling through the cell it can be directly related to the transfer of the ions from one electrode to the other, and this is equivalent to current density (e.g. amp/m2). Hence, the current in the cell is

J = CJiF (it;, ) (17.54)

Fig. 17.4 Ions in cell

If the hydrogen-oxygen cell is taken as an example, it is apparent that the charge is carried across the cell by OH- ions being absorbed on the hydrogen electrode, and being

Page 376: Advanced Thermodynamics for Engineers

360 Fuel cells

produced on the oxygen electrode. These OH - ions are produced in the electrolyte (KOH) as the KOH mvlecule spontaneously dissociates into K’ and OH- ions (in weak aqueous solutions), so ivhile the negative ions travel in one direction the positive ones go in the opposite Ckction. This can be considered schematically as in Fig 17.4, which represents a thin element of the electrolyte. Ions are flowing through the electrolyte, and t+ represents the transport number of the cations, while t - represents that of the anions. If the electrochemical potentials of the two ions on side A of the element are

(17.55)

then it is possible to evaluate the change in Gibbs function, dG, when an infinitesimal number of ions move through the element, giving

(17.56)

Now when the system is in the steady state, dG must be zero, and hence, substituting from eqn (17.55)

(17.57)

-dG= t + d+iu+ - t - d+iOH-

-t+ dpu+ - t + F dV + t - dpon- - t - F d v = O

Since t + + t = 1 then

F dV = - t + dpu+ + t - dpon- (17.58)

Applying Onsager’s relationship, Ji = Cf=, LikX,, gives

(17.59)

The term L,, is equivalent to the conductance of the electrolyte, k , and by the Onsager reciprocal relationship,

t+ k L,, = L , , = +- F

Hence, the current, J , is given by

(17.60)

(17.61)

If the circuit connecting the electrodes is open then no current flows and J = 0, giving the result obtained in eqn (17.58). In the case where a current flows, J>O, then there will be a reduction in the potential available at the electrodes because some of the ‘force’ that can be generated in the cell has to be used to propel the ions through the concentration gradients that exist.

A more comprehensive analysis of the irreversible thermodynamics of the cell would take into account the temperature gradients that might be set up in operating the cell, and

Page 377: Advanced Thermodynamics for Engineers

Problems 361

another group of equations for thermal conduction would have to be added to the equations in the matrix of eqn (17.59).

17.5 Concluding remarks

It has been shown that fuel cells offer the potential of converting the bond energies of a reaction directly to electrial power: this enables the restrictions of the Second Law to be avoided. The theory of fuel cells has been developed from basic thermodynamic principles, and a new property called electrochemical potential has been introduced. Finally, the application of irreversible thermodynamics to the fuel cell has demonstrated where some of its shortcomings will lie.

PROBLEMS Assume the value of the Faraday constant, F = 96 487 kC/kmol.

An electric cell has the following chemical reaction:

Zn(s) + 2AgCl(s) = ZnC1, + 2Ag(s)

and produces an emf of 1.005 V at 25°C and 1.015 V at 0 ° C at a pressure of 1 bar. Estimate the following parameters at 25°C and 1 bar:

(a) (b)

change in enthalpy during the reaction, amount of heat absorbed by the cell, per unit amount of zinc, to maintain the temperature constant during reversible operation.

[ -216 937 kJ/kmol Zn; -23 002 kJ/kmol Zn]

An electric cell is based on the reaction Pb + Hg,Cl,-+PbCl, + 2Hg. If the enthalpy of reaction for this reaction, e,, at 25°C is -95 200 kJ/kg Pb, calculate the emf and rate of change of emf with temperature at constant pressure if the heat transfer to the cell is 8300 kJ/kg Pb.

[0.5363 V; 1.44 x V/K]

Calculate the emf of a hydrogen-oxygen fuel cell operating reversibly if the overall reduction in Gibbs energy is 238 MJ/kmol H,. If the cell operates at 75% of the reversible emf due to internal irreversibilities, calculate the magnitude and direction of the heat transfer with the surroundings, assuming that the enthalpy of reaction at the same conditions is -286 MJ/kmol H,.

[ 1.233 V; -53750 kJ/kg H2]

An ideal, isothermal, reversible fuel cell with reactants of oxygen and hydrogen, and a product of water operates at a temperature of 400 K and a pressure of 1 bar. If the operating temperature increases to 410 K, what must be the new pressure if the open circuit voltage is maintained constant? The values of Kp, are 1.8 134 x lo2' at 400 K and 2.4621 x at 410 K.

[2.029 bar]

Page 378: Advanced Thermodynamics for Engineers

362 Fuel cells

5 A hydrogen-oxygen fuel cell is required to produce a constant voltage and operate over a pressure range of 0.125 to 10 bar. The datum voltage is 1.16 V at a temperature of 350 K and pressure of 10 bar. If all the streams are at the same pressure evaluate the range of temperature required to maintain the voltage at this level, and show how the operating temperature must vary with pressure.

Assume that the enthalpy of reaction of the cell, Qp = -286 000 kJ/kmol H,, and that this remains constant over the full temperature range. The valency, zi = 2.

[317.4 K]

6 A hydrogen-oxygen fuel cell operates at a temperature of 450 K and the reactants and products are all at a pressure of 3 bar. Due to internal resistances the emf of the cell is only 70% of the ideal value. Calculate the ‘fuel consumption’ of the ideal cell and the actual one in g/kW h.

An alternative method of producing electrical power from the hydrogen is to bum it in an internal combustion engine connected to an electrical generator. If the engine has a thermal efficiency of 30% and the generator is 85% efficient, calculate the fuel consumption in this case and compare it with that of the fuel cell. Explain why one is higher than the other.

Table P17.6 Table of values of enthalpy and Gibbs energy at T = 450 K

Species Enthalpy, h (kJ/kmol) Gibbs energy, g (W/kmol)

Oxygen, 0, 13 125 -84 603 Hydrogen, H, 12 791 -51 330 Water, H,O -223 672 -314 796

[32.25 g/kW h; 46 g/kW h; 116.2 g/kW h]

Page 379: Advanced Thermodynamics for Engineers

Bibliography

This bibliography lists some of the references that have been used in preparing the book, and on which some of the ideas were based. In general it is simply a bibliography and not a set of specific references, although some material has been drawn from the material in this list; where that is the case a specific reference is made in the text. There is not a bibliography for all chapters because the information is subsumed into the books relating to adjacent chapters.

Chapter 1 State of equilibrium

Atkins, P W, 1994: The Second Law - energy, chaos, and form. Scientific American

Benson, R S, 1977: Advanced Engineering Thermodynamics. Pergamon Denbigh, K G, 1981: The Principles of Chemical Equilibrium. Cambridge University Press Hatsopoulos, G N and Keenan, J H, 1972: Principles of General Thermodynamics. John

Keenan, J H, 1963: Thermodynamics, John Wiley, New York, 14th printing

Books

Wiley

Chapter 2 Availability and exergy

Ahearn, J E, 1980: The Exergy Method of Energy Systems Analysis. John Wiley Chin, W W, and El-Masri, M A, 1987: Exergy analysis of combined cycles: Part 2 -

analysis and optimisation of two-pressure steam bottoming cycles, Transactions of the American Society of Mechanical Engineers, 109,237 -243

El-Masri, M A, 1987: Exergy analysis of combined cycles: Part I - air-cooled Brayton- cycle gas turbines, Transactions of the American Society of Mechanical Engineers, 109, 228-236

Goodger, E M, 1979: Combustion Calculations. Macmillan Haywood, R W, 1980: Equilibrium Thermodynamics for Engineers and Scientists, John

Heywood, J B, 1988: Internal Combustion Engine Fundamentals. McGraw-Hill Horlock J H, and Haywood, R W, 1985: Thermodynamic availability and its application to

combined heat and power plant, Proceedings of Institution of Mechanical Engineers,

Wiley

199, C1, 11-17

Page 380: Advanced Thermodynamics for Engineers

364 Bibliography

Keenan, J H, 1963: Thermodynamics. New York: John Wiley, 14th printing Kotas, T J, 1995: The Exergy Method of Thermal Plant Analysis. Florida: Kreiger Moran, M J, and Shapiro, H N, 1988: Fundamentals of Engineering Thermodynamics.

Wiley International Edition Patterson, D J, and Van Wylen, G J, 1964: A digital computer simulation for spark-ignited

engine cycles, Society of Automotive Engineers Progress in Technology Series, 7, Society of Automotive Engineers

Rant, Z , 1956: Exergie, ein neues Wort fur ‘technische Arbeitsfahigkeit’ (Exergy, a new word for ‘technical work capacity’), Forsch Gebeite Zngenieurwes, 22, 36

Rogers, G F C, and Mayhew, Y R, 1988 Thermodynamic and transport properties of fluids. SI units; Third edition. Oxford, Blackwell

Tsatsaronis, G and Winfold M, 1985: Exergoeconomic analysis and evaluation of energy- conversion plants: Part I - A new general methodology; Part 11 - Analysis of a coal- fired steam power plant, Energy, 10,69-80, and 81-94

Chapter 3: Pinch technology

Barclay, F J, 1995: Combined Power and Process - an exergy approach. London,

Linnhoff, B, and Senior, P R, 1983: Energy targets clarify scope for better heat

Linnhoff, B, and Turner, J A, 1981: Heat recovery networks: new insights yield big

Smith, R, 1995: Chemical Process Design. McGraw-Hill

Mechanical Engineering Publications

integration. Process Engineering, Mar, 29-33

savings. Chemical Engineering, Nov, 56-70

Chapter 4: Rational efficiency

Haywood, R W, 1975: Analysis of Engineering Cycles. Pergamon Haywood, R W, 1980: Equilibrium Thermodynamics for Engineers and Scientists. John

Horlock, J H, 1987: Co-generation - combined heat andpower. Pergamon Wiley

Chapter 5: Endoreversible engines

Bejan, A, 1988: Advanced Engineering Thermodynamics. Wiley International Edition El-Masri, M A, 1985: On thermodynamics of gas turbine cycles: Part I - Second law

analysis of combined cycles. Transactions of the American Society of Technical Engineers, 107, 880-889

Chapter 6: Relationships between properties

This subject is covered in most of the general texts on engineering thermodynamics. The reader is referred to these to obtain additional information or to see different approaches.

Chapter 7: Equations of state

This subject is covered in most of the general texts on engineering thermodynamics. The reader is referred to these to obtain additional information or to see different approaches.

Page 381: Advanced Thermodynamics for Engineers

Bibliography 365

Chapter 8: Liquefaction of gases

Haywood, R W, 1972: Thermodynamic Tables in SI Units. Cambridge University Press Haywood, R W, 1975: Analysis of Engineering Cycles. Pergamon Haywood, R W, 1980: Equilibrium Thermodynamics for Engineers and Scientists. John

Wiley

Chapter 9: Thermodynamic properties of ideal gases

Benson, R S, 1977: Advanced Engineering Thermodynamics. Pergamon JANAF, 1971 : Thermochemical Tables. Washington, DC: National Bureau of Standards

Publications, NSRDS-N35 37

Chapter I O : Thermodynamics of combustion

Glassman, I, 1986: Combustion. New York: Academic Press Goodger, E M, 1979: Combustion Calculations. Macmillan Strahle, W C, 1993: An Introduction to Combustion. Gordon and Breach

Chapter 11: Chemistry of combustion

Benson, R S, 1977: Advanced Engineering Thermodynamics. Pergamon Benson, S W, and Buss, J H, 1958: Additivity rules for the estimation of molecular

properties, and thermodynamic properties. J . Chem Phys, 29, No 3 Goodger, E M, 1979: Combustion Calculations. Macmillan Kuo, K K, 1986: Principles of Combustion. Wiley International Edition

Chapter 12: Chemical equilibrium and dissociation

Benson, R S, Annand, W J D, and Baruah P C, 1975: A simulation model including intake and exhaust systems for a single cylinder four-stroke cycle spark ignition engine. International Journal of Mechanical Sciences, 17,97- 124

Denbigh, K G, 1981: The Principles of Chemical Equilibrium. Cambridge: Cambridge University Press

Horlock, J H, and Winterbone, D E, 1986: The Thermodynamics and Gas Dynamics of Internal Combustion Engines, vol II. Oxford: Oxford University Press

Heywood, J B, 1988: Internal Combustion Engine Fundamentals. McGraw-Hill Lavoie, G A, Heywood, J B, and Keck, J C, 1970: Experimental and theoretical study of

nitric oxide formation in internal combustion engines, Combustion Science and Technology, 1,313-326

Vickland, C W, Strange, F M, Bell, R A, and Starkman, E S A, 1962: A consideration of high temperature thermodynamics of internal combustion engines. Society of Automotive Engineers Transactions, 70, 785-795

Chapter 14: Chemical kinetics

Annand, W J D, 1974: Effects of simplifying kinetic assumptions in calculating nitric oxide formation in spark ignition engines. Proceedings of Institution of Mechanical Engineers, 188, (41/74), 431 -436

Page 382: Advanced Thermodynamics for Engineers

366 Bibliography

Daneshyar, H, and Watfa, M, 1974: Predicting nitric oxide and carbon monoxide concentrations in spark ignition engines. Proceedings of Institution of Mechanical Engineers, 188, (41/74), 437-443

Glassman, I, 1986: Combustion. New York: Academic Press Heywood, J B, 1988: Internal Combustion Engine Fundamentals. McGraw-Hill Heywood, J B, Faye, J A, and Linden, L H, 1971: Jet aircraft air pollutant production and

bavoie, G A, Heywood, J B, and Keck, J C, 1970: Experimental and theoretical study of hitric oxide formation in internal combustion engines. Combustion Science &

dispersion. AZAA Journal, 9 ( 5 ) , 841 -850

Technology, 1, 313-326

Chapter 15: Combustion and flames

Abdel-Gayed, R G, Bradley, D, and Laws, M, 1987: Turbulent burning velocities: a general correlation in terms of straining rates. Proceedings Royal Society, London,

Bradley, D, Lau, A K, and Lawes, M, 1992: Flame stretch rate as a determinant of turbulent burning velocity. Philosophical Transactions Royal Society, London, A338,

Damkohler, Z, 1940: Elektrochem, 46, 601 (Translated as NACA Technical Memo 11 12

Gaydon, A G, and Wolfhard, H G, 1978: Flames: their structure, radiation and

Glassman, I, 1986: Combustion. New York: Academic Press Heikal, M R, Benson R S, and Annand W J D, 1979: A model for turbulent burning speed

in spark-ignition engines. Proceedings of Conference of Institution of Mechanical Engineers on Fuel Economy and Emissions of Lean-burn engines, London

Heywood, J B, 1988: Internal Combustion Engine Fundamentals. McGraw-Hill Kuehl D K, 1962: Laminar burning velocities in propane-air mixtures. Eighth Inter-

Kuo, K K, 1986: Principles of combustion. Wiley International Edition Lancaster, D R, Kreiger, R B, Sorenson, S C, and Hull W, 1976: Effects of turbulence on

spark-ignition engine combustion. Society of Automotive Engineers, 760160, Detroit Lewis, B, and von Elbe, G, 1961: Combustion, Flames and Explosions of Gases, 2nd edn.

New York: Academic Press Mallard, E and le Chatelier, H L, 1883: Ann Mines, 4, 379 Metgalchi, M, and Keck, J C, 1980: Laminar burning velocity of propane-air mixtures at

high temperature and pressure. Combustion & Flame, 38, 143-154 Metgalchi, M, and Keck, J C, 1982: Burning velocities of mixtures of air with methanol

iso-octane, and indolene at high pressure and temperature. Combustion & Flame, 48,

A414,389-413

359 - 387

- The effects of turbulence on flame velocities in gas mixtures)

temperature. London: Chapman & Hall

national Symposium on Combustion

191 -210 Strahle, W C, 1993: An Introduction to Combustion. Gordon and Breach Whitehouse, N D, and Way, R J B, 1970: Rate of heat release in diesel engines and its

correlation with fuel injection data. Proceedings of Institution of _Mechanical Engineers, 184, 3J, 17-27

Zel’dovitch, Y B, 1948: J Phys Chem USSR, 22 (1)

Page 383: Advanced Thermodynamics for Engineers

Bibliography 367

Zel’dovitch, Y B, and Frank-Kamenetsky, D A, 1938: Compt Rend Acad Sci USSR, 19,

Zel’dovitch, Y B, and Semenov, N, 1940: Journal of Experimental and Theoretical 693

Physics USSR, 10, 11 16 (Translated as NACA Tech Memo 1084)

Chapter 16: Irreversible thermodynamics

Bejan, A, 1988: Advanced Engineering Thermodynamics, Wiley International Edition Benson, R S, 1977: Advanced Engineering Thermodynamics. Pergamon de Groot, S R, 1951: Thermodynamics of Irreversible Processes. North Holland Denbigh, K G, 1951: Thermodynamics of the Steady State Methuens Monographs on

Fick, A, 1856: Die medizinische Physik. Hoase, R: Thermodynamics of Irreversible Processes. Addison-Wesley Lee, J F, and Sears, F W, 1963: Thermodynamics. Addison-Wesley Onsager, L, 1931a: Reciprocal relations in irreversible processes, part I. Phys Rev, 37,

Onsager, L, 1931b: Reciprocal relations in irreversible processes, part 11. Phys Rev, 38,

Tribus, M, 1961 : Thermodynamics and Thermostatitics. Van Nostrand Yougrau, W, Merwe, A, and Raw, R, 1966: Treatise on Irreversible and Statistical

Chemical Subjects

405-426

2265-2279

Thermophysics. Macmillan

Chapter 17: Fuel cells Berger, C (Ed), 1968: Handbook of Fuel Cell Technology. Prentice-Hall Denbigh, K G, 1981: The Principles of Chemical Equilibrium. Cambridge University

Hart, A B, and Womack, G J, 1967: Fuel Cells. London: Chapman and Hall Haywood, R W, 1980: Equilibrium Thermodynamics for Engineers and Scientists. John

Liebhafsky, H A, and Cairns, E J, 1968: Fuel Cells and Fuel Batteries. Wiley

Press

Wiley

Page 384: Advanced Thermodynamics for Engineers

Index

Index of tables of properties

Table 8.1 Table 8.2

Table 9.1

Table 9.2 Table 9.3 Table 9.4

Table 10.2 Table 11.1 Table 11.2 Table 12.3 Table 15.1

Critical temperatures and pressures for common substances Properties of air on saturated liquid and vapour lines, and superheated air at low temperatures and high pressures Molecular weights of elements and compounds commonly encountered in combustion Values of the universal gas constant, %, in various units Enthalpy coefficients for selected gases found in combustion processes Properties of gases Oxygen, nitrogen Carbon monoxide, carbon dioxide Hydrogen, water Nitric oxide, methane Enthalpies of formation of some common elements and compounds Some atomisation, dissociation and resonance energies Enthalpies of formation and reaction for commonly encountered fuels and compounds Equilibrium constants Flammability and explosion limits at ambient temperature and pressure

135

152

159 160 164

167 168 169 170 192 209 217 257-8 296

acetvlene 185.217.300 ammonia 296 . . acidrain 286 activation energy 279-85, 31 1 activity coefficient 353 adiabatic combustion 187-205 adiabatic compressibility 115 adsorption, gas molecules 294 air 152, 159

air standard cycle 71, 189,208 ai-fuel ratio 30-1, 186-8, 225, 265-72, 292

lean(weak) 186,223,245,265-72.301 rich 186,222,224,245,252,265-72,301 stoichiometric 186,216-7,233,267,269,299

composition 159

alcohols 185, 213,217 alkanes 138, 185, 217 alkenes 185, 217 Amagat’s Law 121 -2

amount of substance 7, 100, 158-9, 172-7, 189,

annular combustion chamber 3 13 aqueous solution 347-53 argon 159 aromatics 217 Arrhenius equation 279,294,299,311 atmospheric nitrogen 159 atomic nitrogen 242-5.273.282 atomic oxygen 164,242-5,273,282-4 atomic weight 158 atomisation 21 1

energy 208-15 availability 13-36, 64, 359

balance, 27-36 change of 18.22-9 closed system 18, 27-30

196-205,219-59,268,275, 319, 348

Page 385: Advanced Thermodynamics for Engineers

370 Index

availability (cont.) effect of change of entropy 32 effect of change of volume 30 effect of combustion 30-4 flow 21.35 loss of 13-36 molar 15 non-flow 15, 18,30 open system 34-6 rate of change 29,35 reaction 31,40-1 specific 15, 19,27-9 steady flow 21,35 work 29

availability transfer 29 available energy 21-6,36,67 Avogadro’s

hypothesis 159 number 340, 347

barrel swirl (tumble) 307 Baruah 268 Beattie-Bridgeman equation 128 Bejan 85 Benson 164 benzene 202.209-10,217,225 benzoic acid 2 15 Bertholet equation 129 binary cascade cycle 137 boiler 276, 29 1 boiling point 137 Boltzmann constant 340 bond energy 40, 182, 187,208-17,291,348

atomisation energy 209- 17 dissociation energy 209- 17 minimum potential energy 208- 11

Bosch Smoke Number 287 Boyle’s Law 121-2 Bradley 303 brake thermal efficiency 304 branching reactions 295-6 bronchial irritants 286 Bunsen burner 297,305 butane 135-6, 185, 192,217,296

calorific value 64 higher 188-205 lower 188-205

carbon-I2 158-9, 192 carbon dioxide 9,40, 135, 139-40, 159, 164, 168,

carbon monoxide 1.9.40, 159, 164, 168, 184-92,

carbon particulates 285,287,312 carbon/hydrogen ratio 185,224,270,272 Carnot

182-205, 209,211,218-75,292, 345

218-75,285,288,296,300

cycle 21, 65, 68, 85, 89

efficiency 65-82,85-96,345

binary cycle 137 ternary cycle 137

cascade 137

catalytic converters 288 cetane 217

chain reaction 295-6 Charles’ Law 121-2 chemical bonds 187,208-17,354 chemical composition 186-8, 196-206,229-38,

chemical equilibrium 218-57 chemical kinetics 8, 194,245,265,275,276-89,

chemical potential 7.8, 100, 165,220-31,319,332,

number 3 11

245-57,265-85

295,299,306

334-7.349-53 Standard 227-30

chemical reaction 185-285 chemical structure of fuels 208-17 chemistry, combustion 184-7,208-17 chlorine 351 Clapeymn equation 117 Clausius equation of state 123 Clausius inequality 14 Clausius-Clapeyron equation 115. 117 closed system 5-7, 13-21,27-34,218 coal 185 coefficient of expansion 114, 123-5, 144 coefficient of performance 137-8 cold utility 47-60 combined cycle gas turbine (CCGT) 47,92,287 combined cycle heat engine 92-6, 137 combustion 182-313

adiabatic 187-274 chemistry 184-7,208-17 constant pressure 193,202,312 constant volume 193, 196, 199-202,233-8,

diffusion 291 energy equation 188-205.245-57 exergy change 40 heat release 208,312 heterogeneous 183,292,305-6 homogeneous 183,292,296-305 incomplete 195,204,233-8,245-57 initiation 304-5.310

multi-phase 183 non-premixed 183,305-6,309-13 premixed 292-305,307-9 rich mixture 186, 195,202-5,224,233-8.252-4,

single-phase 183

three dimensional 183 turbulent 183, 302-3

245-57,267-75

lam in^ 183,296-302

267-74

Speed of 297 - 304

Page 386: Advanced Thermodynamics for Engineers

Index 371

two-dimensional 183 weak mixture 186,200,223,245-52.267-74 with heat transfer 194, 199 with work transfer 194 zero-dimensional 183

combustion chamber 307-13 combustion systems

diesel 309- 12 gas turbine 312-3 spark ignition 307-9

combustion zones, gas turbine 313 composite diagram 5 5 , 6 0 composite temperature - heat load diagram 5 1,57 compressibility

adiabatic 115 factor 129 isentropic 115 isothermal 114-5,122,125

coefficient 129 isentropic 15-8,24-5.65-82

compression

compressor 24-5.76-82 concentration

component 3 19 gradient 316,332,335,360 molar27744

condenser 69-76 conductance

conduction heat transfer 92-5

of electricity 318,322-32 of heat 317,319-32

fluxes 319-42 forces 319-42

sulfate 347-51

con jugate

copper 347-51

corresponding states, law of 125-9 Coulombic forces 348 counterflow heat exchanger 37-40 coupled

equations 316-42 processes 319-42

coupling matrix 319-42 critical

isotherm 125-7

pressure 124, 135, 140 temperature 124, 135, 146 volume 124

water 126, 135

point 124-31, 135

Critical point 124-31.135

cross-coupling 316-42 crude oil 184 current

electric 316.322-32.347-61 density 321-32

cycle calculations 30-4.48, 69-82 cycle efficiency 69-82 cycles

air standard 71, 189,208 bottoming 34 Camot 21,65,68, 85, 89 Diesel 68 endo-reversible 65-82, 85-96 gas turbine, see Joule internally reversible 65-82, 85-96 irreversible 65-82 Joule 68.76-82

Rankine 69-76,79 reversible 65 Stirling 85

Otto 30-4,68,267

cyclone separator 288

Dalton Principle 172 Damkohler 302 Daniel1 cell 345 -5 1 datum pressure, also standard pressure 164,228-9 dead state 14-43,65-82

temperature14-43.65-82 pressure 14-43

Debye theory 162 deflagration 183,296-305 de Groot 333 delocalised electrons 210 desorption

destruction of availability 24-9, 39-41 detonation 183, 296, 309 diatomic 184 Diesel

gas molecules 294

engine 309-12 cycle 68 oil 217

Dieterici equation of state 129 diffuser 313 diffusion 175,294, 319, 332-39, 359

burning 287.305-6,311 flames 291,297, 305-6

mass 298,332-42 thermal 298,303,332-42

dilution zone 313 direct injection engines 309-12 dissociation 7, 182, 194-5.218-59, 265-75

diffusivity

degree of 218-59 effect of pressure 239-40 effect of temperature 239-41

dissociation energy208-10, 214 dry ice (carbon dioxide) 139 dryness fraction (quality) 137, 150 Dufour effect 332-8 dynamic equilibrium 218,222, 349

Page 387: Advanced Thermodynamics for Engineers

372 Index

eddy diffusivity 302 efficiency

at maximum power 85-96 cycle 64-82 fuel cell 357-9 isentropic 22-5,30-4,69-82 overall 64 rational 65-82, 154, 359 thermal 17,64-82,85-96,292,355

cells 346-51 conduction 316,323-32,345-61 flow 323-32

electrochemical potential 349-61 electrode 346-51 electrostatic precipitator 288 elementary reaction 278 end gas 309 endoreversible heat engine 85-96 endothermic reaction 184,256,273-5,295 energy

electrical

available 13-36 Gibbs see Gibbs energy Helmholtz see Helmholtz energy internal 5-9, 13-34,37-40, 100-18, 160-78,

potential 1-4 unavailable 6.21-2.26, 37, 67, 80

187-205.245-57

energy equation, 5-6, 14-5,27, 34.58, 86, 143, 189-205,233-57,347-9

steady flow, 58,89, 143, 189-205 unsteady flow 34

combustion systems 307 - 13 diesel 276,286-7,291,305, 309-12 irreversible 65 petrol see spark-ignition reversible 5, 17,21,26,65, 85-96 spark-ignition 208,276,291, 304, 307-9 stratified charge 305

engine

enthalpy 6-9,22-5.34-6,40-3, 100-18, 141-54, 160-75, 187-205,245-57

molar 160, 165, 174 of formation 65, 187-205,209-17,241-2 of products 188-205.245-57 of reactants 188-205, 245-57 ofreaction 40-1, 64, 187-217, 355-61 specific 160, 337

enthalpy coefficients of gases 167-70 enthalpy-temperature diagram 188 -205 entropy 3-7, 19-43,64-82.91, 100-18, 162-78,

218-21,316-42 change 3, 19,23, 28,37,39, 117, 348 change due to mixing 1754,229-31 creation 3-4, 317-22, 333, 336 flow 5,317-22 flux 321 -36

generation, see creation ideal gas 23-5.77-82, 162, 165-6 maximum 4,3 19 mixtures 175-8,229-31

production, see creation environment, see also dead state 23 equation of constraint 227 equations of state 121-31, 158-9

Beattie-Bridgeman 128 Bertholet 129 Dieterici 129 idealgas 105-7,110, 121-3, 144, 159-78,

perfect gas, see also ideal gas 15-9,30-4,

van der Waals’ gas 123-8 virial 128

35 1-8

of transport 324-32

195-205,233-8,245-57

76-82

equilibrium 1-9,208,218-57.265-6, 317, 345,

chemical 218-57,351-8 condition 1-9,275-6,282 dynamic 218,222,349 mechanical 1-2, 14 metastable 2,8,211,292 neutral 14 of mixtures 225-33 stable 1-4, 8 static 222, 317,333 thermal 1-9 thermodynamics 1-9,345 unstable 1-4, 8, 140

equilibrium constant 228-59,283,354 equivalence ratio 186-205,222-5,233-57,

ethane 135, 185, 192,217,296 ethanol 185, 192,213,217 evaporation 115-7, 124 exergy 13.36-43,64-82,359

296-306

change 37-9,42 of reaction 40-1

combustion 182-3 13 reaction 188-90

isenthalpic 141 -54 isentropic 22-4, 30-4, 69-82 non-isentropic 22-4,69-82

exothermic

expansion

expansion, coefficient of 114, 123, 125, 144 explosion limits 294-6 external irreversibility 68, 73, 85-96

Faraday’s law of electrolysis 347 Faraday constant 347-60 Fick’s Law 319,332 First Law of Thermodynamics 5, 13-5,37,52, 141,

188-205.233-8,245-57,347-8

Page 388: Advanced Thermodynamics for Engineers

Index 373

Fischer-Tropsch process 188 flame 291-313

diffusion 291,297, 305-6 extinction 303-4 laminar 183,296-302 premixed 292-305,307-9 stretch 303-4

flame quenching 303-4.308 flame speed 292-303

turbulent 302-3 laminar 292-302

flame speed factor 302-3 flame traverses charge 291 flammability limits 296, 303-4 flash tank 139 flue gas desulfurisation 286 fluorine 293 Fourier’s law of conduction 316-32 fraction of reaction 226-31 Frank-Kamenetskii 297 fuel 184-92,208-17

structure 208- 17 fuel cell 289, 345-61

hydrogen-oxygen 345,353,356,359 fuel injection 291, 306, 310 fuel-air ratio, see air-fuel ratio functional relationships 100- 18

galvanic cell 345, 350 gas

ideal 105-7, 110, 121-3, 144, 158-78, 195-205,

perfect see also ideal gas 15-9, 30-4,76-82 superheated 75 -6, ,101, 1 1 5 , 1 2 1 - 8 van der Waals’ 123-8

specific 113, 125-9, 158 universal 123-8, 160

233-38.245-57

gas constant

gasoline 2 17 gas turbine 76-82,276,287,291, 305-6, 312-3 Gay-Lussac’s law 121 -2 Gibbs 36 Gibbs energy 6-9, 13-4,65, 124, 129, 130-1,

163-6,218-33,265,289,292,345,351-61 effect of composition 218-21, 229-31 effect of pressure 219 effect of temperature 219 minimum 8,231,292-3 molar 163-5,352 specific 165

Gibbs function see Gibbs energy Gibbs potential see Gibbs energy Gibbs-Dalton laws 172-5 Glassman 286 global warming 287 governing

qualitative 292

quantitative 292 greenhouse effect 287

Hartley 319,332 Hatsopoulos 2 Haywood 15,42, 152 heat

flow 317-39 of formation 65, 187-205,208-17,241-2 of reaction 64, 187-217,238-9,355-59 of transport 326,331,338

heat cascading 57 heat engine

endo-reversible (internally reversible) 85 -96 externally irreversible 85-96 irreversible 64-82 reversible 5-7, 14-21,64-82, 85-96

counterflow 37-40 irreversible 39-40, 154 reversible 37-9

capacity 49-60 rate 321-2

diagram 3 12

coefficient 86 conductance 88-92 irreversible 39-40,85-96 network 49-60 rate 24, 86-92, 316-41 resistance 90 reversible 5, 15, 37-9, 220

heat exchanger 37-40, 150-4

heat flow

heat release (combustion) 312

heat transfer 37-40,47-60,85-96,321-39

heavy fuel 184, 286 Heikal302 Helmholtz

energy 5-9, 13-4 function see Helmholtz energy

heptane 185,217 Hess’ law 190-1,204,208,235,245-57 heterogeneous combustion 183,291, 305-6 hexane 217 Heywood 30,245,282-3 homogeneous combustion 183,292,305-6 Honda 292 Horlock 245, 268 hot utility 47-60 hydrocarbons 40-1, 136-8, 159, 182-7,208-17,

224-5,245-57,265-75,288,345 saturated 185 unsaturated 185

hydrochloric acid 351 hydrogen 135, 148, 164, 169, 192.293-6.300,

351-9 engine 288 oxygen fuel cell 345-59

Page 389: Advanced Thermodynamics for Engineers

374 Index

hypergolic mixture 293-4.31 1

ideal gas see gas, ideal

ignition 291-3,304,307.310-1 mixtures see mixtures of ideal gases

delay 3 11 minimum energy 305 timing 303

indirect injection diesel engine 309- 10 injection

direct 309-10 fuel 309 indirect 309 nozzle 309

intercooled regenerated gas turbine 287-8 interface 346-50 intermediate combustion zone 313 internal combustion engine see engine internal energy, 5-6, 18-20,26-34,37-40, 100-5,

109-11, 160-2, 167-70,174-5,187-205, 233-57,347-9

molar, 160,165,176 of formation, 187-8,208,212 of reaction, 187-9,235,248-9

reactants, 189-205.233-57 specific, 100, 160, 165

products, 189-205,233-57

internal energy-temperature diagram 190, 194,265 internally reversible 21 interval heat balance 52 inversion point 142, 143, 147,150 inversion temperature 142

maximum 145-6 minimum 145-6

irreversibility 21-36.65-82, 154 irreversible

engine 65 -82 heat transfer 3,39-40,85-96 thermodynamics 9,316-41

isenthalpic line 142-54 isenthalpic process 142-54 isentropic

compression 15-8,24-5,65-82 expansion 22-4,30-4.69-82

isentropic compressibility 1 15 isentropic efficiency 22-5.30-4.69-82 iso-octane 192,217,300 isobar 115-8, 129-31 isotherm 115-8,123-31

isothermal compressibility 114-5, 122, 125

jets 305-6,310 Joule experiment 110 Joulean heating 323-34 Joule - Thomson

critical 125-31

coefficient 106, 141-2, 151

effect 143-8 experiment 141-9

junction. thermocouple 322-32

Karlovitz number 303-4 Keck 299 Keenan 2,4,36 kinematic viscosity 302 kinetic theory of gases 123,278,316 kinetics

chemical, see chemical kinetics rate, see chemical kinetics

knock 183,296,309 Knudsen 339 Kuehl301

laminar flame speed 296-305 theories 297-9 thickness 303

Lancaster 303 latent heat 115-8.210-7 Lavoie 245,281-2 law of corresponding states 125-6 law of mass action 225-32.278 Laws of Thermodynamics

First 5,13-5,37,52. 141,188-205.233-8, 245-57.347-8

Second 2-9.13-43.64-82.100-18.175-8 Third 162-3

Le Chatelier 297 lean burn engines 304 lean mixture, see air-fuel ratio Lewis 299

Linde liquefaction plant 148-54 Linnhoff 47,49 liquefaction 135-54 liquefied natural gas 136,138 liquefied petroleum gas (LFG) 136 liquid line

liquid phase 113-8,123-31,135-54,210-7,346-58 luminous region 297 Lurgi process 188

macroscopic theories 3 16 Mallard 297 mass action, law of 225-32,278 mass m s f e r 320,332-42 maximum power 85-96 maximum work 5-6,15-36.73

Number 298,303-4

saturated 113-8,123-31, 135-54

Useful 6.15-36,64-79,351,357 output 85-96

Maxwell equal area rule 129-31 relationships 100-18, 144

Page 390: Advanced Thermodynamics for Engineers

Index 375

May combustion chamber 307-8 Metgalchi 299 methane41, 135, 164, 170, 185,191-2,196,212.

217,240-2,256,267-75,296,300 exergy of reaction 4 1

methanol 185, 192,217,300,345 minimum energy for ignition 305 minimum net heat supply 47-60 minimum temperature difference 47 -60 misfire 304 mixing zone, flame 306 mixture 185-7,222-5,265,276

air-fuel ratio 30-1,186-9,225,265-75.292 entropy 175-8 fuel-air ratio, see air-fuel ratio hypergolic 293-4,311 ideal gases 158, 165, 172-8,276

specific heats 175 stoichiometric 186,216-7,233,267,269,299 strength, see air-fuel ratio weak or lean 186,222,245,265-71,300

availability 15 concentration 277-84 properties 160, 165, 174

rich 186,222,224,233-8,245-54,265-75.300

molar

mole fraction 173,229-30,239,245,268,270.

molecular weight 158-9,217,299 Mollier (h-s) diagram 150 multiplication factor (combustion) 295-6

naphthalene 209,217 natural gas 137-8, 185,217,272

liquefaction 137-8 net rate of formation 276-85 net work 19.64-82,348 network heat transfer 49-60 nitric oxide 164, 170, 182,225,242-7.254-7,

273-4,283-5

273-5,280-9 formation 182,242,257,273-5,280-9 initial rate of formation 282-5 prompt 280 thermal 280

atmospheric 159 atomic 164,242-5,273-5,282 oxides of 266,273,276,286,345

nitrogen 135,145, 148, 159, 164, 167,257,270

nitrous oxide 282 NO see nitric oxide non-available energy 6,21-2, 37 non-flow availability 15-8,30 non-isentropic expansion 22-4,69-82 NO, see oxides of nitrogen number density 277

Octane 185,217,245-54,267-75.300

enthalpy of reaction 32,217 exergy of reaction 41

Ohm’s law 316-8,324 Ohmic heating 329 Onsager 3 16

optimum temperature ratio 87-96 orimulsion 286 Ostwald 345 Otto cycle 30-4,68,267 oxidant 182 oxides of nitrogen see nitrogen, oxides of oxygen9, 135,145, 148, 159,164,167, 185, 192,

OH-radical243-5,281,293-5

reciprocal relationship 319-22.342.360

196-205.208-17.218-57,265-75, 292-306

atomic 164,242-5,273-5,282-5 ozone 286

p-v-T diagram 8, 101, 105 PAH (polycyclic aromatic hydrocarbons) 287 PAN (peroxyacyl nitrate) 286 paraffins 185,217,272

particulate emissions 287,312 Peltier

partial PESSW 173, 176,229-33,246,257-8,312

coefficient 322-32 effect 316-32

pentane 217 perfect gas 15-9,304,76-82

flow exergy 42-3,78 law 121-3

petrol 217,300 engine see spark ignition

boundary 346 liquid 113-8.123-31, 135-54,210-7, 346-58 supercritical 140 superheat 123-9, 135, 140-54

phase 101-18

phenomenological laws 316 photochemical smog 286 pinch

point 47-60 technology 47-60

piston speed mean 302-3.311

pollutants 187,265,276,286-9 porous plug 142,339-41 potassium hydroxide 353 potential

chemical 7-8, 100, 165,220-31,319,332-7,

difference 321,327,347,352 electro-chemical349-61 energy 1-4 core, flame 306 gradient 316-32

349-53

Page 391: Advanced Thermodynamics for Engineers

376 Index

power generation plant 47,64 power output, maximum 85-96 F'randtl number 303 pre-exponential factor 279, 299 premixed combustion 291-305, 307-9, 311 premixed flames 296-305 preparation rate 3 1 1 -2 pressure

critical 124, 135, 140 datum 164,227-8 partial 172-8,227-59.283-5,353-8 standard 227 - 8

primary combustion zone 3 13 process

integration 47-60 irreversible 3,23-41,65-82, 85-96, 150-4,

isentropic 15-8,22-5,304, 65-82 isobaric 129-131,312-3 isothermal 115-8, 129-31

reversible 23-41, 65-82

316-42

plant 47-60

products temperature 192-205,233-59 propane 135, 136, 185, 192,200-2, 217,296,

300 propanol 217 propyne 2 17 Property

quasi- 36 pseudo- 36 reduced 125-6

proton charge 347 pyrolysis 312

quality (dryness fraction) 137 quenching, flame 187,303,308 quiescent combustion chamber 310

radicals 244-5, 281,295 Rankine cycle 69-76,220 Rant 36 rate constant see reaction, rate constant rate kinetics 8, 194,245,265,275,276-89,295,

rate 299, 306

of entropy production 317-42 of formation 278-8 of heat transfer 2-4, 86-92, 316-42 of reaction 182,276-85

ratio of specific heat capacities 114-5 rational efficiency 64-82, 154, 359 reactants 8.41, 182-205,208-17,222-59,

reaction 182-300 291-313,348

availability of 3 1, 40- 1 backward 276-85 branching 295-6

chemical 186-285 enthalpy of 64, 187-217,238-9.355-61 exergy of 40- 1 forward 276-85 fraction of 226-30 internal energy of 30-4.40-1, 187-217,233-8,

order 277,293-4 rate 182,276-85,295-6,311-2 rate constant 278-85 rate controlled 279-85, 295-6

pressure 125-9, 148 properties 125-9 temperaturel25-9, 148 volume 125-9

refrigeration cycle 135-8 refrigeration plant 137-9 refrigerator 136, 148 relationships

relative molecular mass see molecular weight relaxation time 282 reservoir 14, 17,52,86-96 resonance energy 209- 11 reversible

245 - 57

reduced

thermodynamic 100-18

engine 5, 17,21,26,65,85-96 heat transfer 5, 15,37-9,220 internally 5 process 15,21, 100 work 15

Reynolds number 303 Ricardo Comet V 309 rich mixture 186,222,224,233,244,252,265-75,

300

saturated liquidline 101, 113-8, 123-31, 135-54 liquid-vapour region 113, 123-31, 135-54 vapourline 101, 113-8, 123-31, 135-54, 158

efficiency 65-82, 154, 289,345 of Thermodynamics 2-9, 13-43,64-82, 100-18

secondary combustion zone 309,313 Seebeck effect 316-32 Semenov 297 Senior 47 shaft work 14 solid-gas region 140 soot 287

Soret effect 332-38 spark ignition engines 208,233,266, 304 specific

Second Law

dry 287

availability see availability internal energy see internal energy enthalpy see enthalpy

Page 392: Advanced Thermodynamics for Engineers

Index 377

exergy see exergy Gibbs energy see Gibbs energy

specific gas constant 126, 158 specific heat capacity 104-18, 122, 160-75

at constant pressure 106-7, 113,298 at constant volume 105, 113 mean 171-2 of solid 162 ratio of 114-5 relationships 104-18

spontaneous change 4 squish motion 307 stability criterion 4 stable equilibrium 1-9 standard

chemical potential 227 -3 1 pressure 227-8 temperature41, 189, 193

dead 14-43.65-82 equations 121, 158-9 transitory 8

State

statistical thermodynamics 165, 175, 316, 340-2 steady flow availability 21-35 steady state 3 17 Steam

plant cycle 47-8, 69-76 turbine 69-76,85

Stirling engine 85 stoichiometric 185-205,217,223-59.267-75,

air-fuel ratio 186,216-7,233,267,269,299 coefficients 221-2,226-8,277-9,354 mixture 186.216-7,222,233,254,288

291-313

stoichiometry 186,221-2,227 sulfur dioxide 285-6 supercritical vapour 140, 149 superheated gas 135, 140, 158 supply temperature 47 -60 swirl 310-3

tumble 307-8 barrel 307-8

closed 13-9, 27-34 constant composition 100- 18 isolated 3-4, 317 multi-component 218-59.265-85, 332-42 open 34-6 single component 100- 18

system

T-sdiagram 16-43,64-81,104,116,135-54 target temperature 47-60 Taylor microscale 303 Tds relationships 108 - 1 1 temperature

adjusted 47-60 critical 124, 135, 146

difference, minimum 47-60 intervals 47-60 inversion 142, 145-6 products 193-205,233-59 ratio 87-96

reactants 193-205,233-59 standard 41, 189, 194-5

target 47-60

optimum 87 -96

supply 47-60

temperature-heat load diagram, composite 47-60 temperature-enthalpy transfer diagram 47 -60 thermal

conductivity 298 diffusion 332-42 efficiency 17,64-82,85-96, 292,355 transpiration 340-2

thermocouple 322-332.350 Cu-Ni 328 Platinum-rhodium 328

thermodynamic properties ideal gas 158-78 ideal gas mixtures 158, 172-8

flow 318 fluxes 316-42,359-60 forces 316-42, 359-60 relationships 100- 18 velocities 3 18

thermodynamics classical 316 irreversible 316-42,345,349,352,359-60 steady state 316-42,359

thermodynamic

thermoelectricity 316-32 thermoelectric

effect 322-32 phenomena 322-32

thermostatics 3 16 Third Law of Thermodynamics 162-3 Thomson 3 16

coefficient 323, 332 heat 323,330- 1

throttle 154 throttling 136 toluene 2 17 transitory state 8 triatomic 190 triple point 20

tumble swirl 307-8 turbine 22-4.47-8, 64-82,312-3

work 22-4,64-82 turbocharger 34 turbulence 291, 302-3

intensity 303 turbulent

entrainment 305

pressure 140

Page 393: Advanced Thermodynamics for Engineers

378 Index

turbulent (conr.) flame speed 302-3 mixing 305-10.312-3

Turner 47.49 two-phase region 115-8, 124-31, 135-54 two-property rule 108, 122

unavailable energy 6,21-2,26,67,80 unburned hydrocarbons (uHC) 288,304,307 universal gas constant 123-8, 160 unsteady flow energy equation (USFEE)

utility 47-60 34-35

Cold 47-60 hot 47-60

valency 347-9,354-6 van der Waals’ gas 105-10, 123-9, 144-6 van? Hoff equation 165,238-9,355 vapour

line, saturated 113-8, 123-31, 135-54 supercritical 140,149 wet 136

virial equation of state 128 volumetric strain 122-3 von Elbe 299 water20,41, 125-31, 135,150, 159, 164, 169, 182,

192,211,270,292,353-9 critical point 126,135 exergy of reaction 41

latent heat 115-8 van der Waals’ equation 125-31

water-gas reaction 232,252,257-8 Way 311 weak mixture 186,223,245,265-71 Weber 339 Whitehouse 3 11 Wilson cloud chamber 8 Winterbone 245,268 work

displacement 13-4 feed pump 72-6 maximum 5-6,15,73 maximum useful 15-36.38.64-79,348-51,356 net 19,6442.347-8 output rate 85-96 reversible 13-42 shaft 14 turbine 22-4.64-82 useful 13-25

xylene 217

Young’s modulus 123

Zel’dovitch 297-8 Zero emissions vehicle (ZEV) 288 zinc 347,349

sulfate 347